dynamic accident sequence analysis using dynamic flowgraph

219
Dynamic Accident Sequence Analysis using Dynamic Flowgraph Method and Markov/Cell-to-Cell Mapping Technique by Chireuding Zeliang A Thesis submitted in Partial Fulfilment of the Requirements for the Degree of Master of Applied Science in Nuclear Engineering The Faculty of Energy Systems and Nuclear Science University of Ontario Institute of Technology 2000 Simcoe Street North, Oshawa, Ontario, Canada, L1H 7K4 July 2018 © Chireuding Zeliang, 2018

Upload: others

Post on 09-Dec-2021

10 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

Dynamic Accident Sequence Analysis

using Dynamic Flowgraph Method

and Markov/Cell-to-Cell Mapping Technique

by

Chireuding Zeliang

A Thesis submitted in Partial Fulfilment

of the Requirements for the Degree of

Master of Applied Science in Nuclear Engineering

The Faculty of Energy Systems and Nuclear Science

University of Ontario Institute of Technology

2000 Simcoe Street North, Oshawa, Ontario, Canada, L1H 7K4

July 2018

© Chireuding Zeliang, 2018

Page 2: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

ii

THESIS EXAMINATION INFORMATION

Submitted by: Chireuding Zeliang

Master of Applied Science in Nuclear Engineering

Thesis Title:

Dynamic Accident Sequence Analysis using Dynamic Flowgraph Method and Markov/Cell-to-

cell Mapping Technique

An oral defense of this thesis took place on July 24, 2018 in front of the following

examining committee:

Examining Committee:

Chair of Examining Committee

Prof. Eleodor Nichita

Research Supervisor

Prof. Akira Tokuhiro

Examining Committee Member

Prof. Glenn Harvel

External Examiner

Dr. Salam K. Ali, Ontario Power Generation

The above committee determined that the thesis is acceptable in form and content and that

a satisfactory knowledge of the field covered by the thesis was demonstrated by the

candidate during an oral examination. A signed copy of the Certificate of Approval is

available from the School of Graduate and Postdoctoral Studies.

Page 3: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

iii

ABSTRACT

In the recent years, numerous concerns have been raised regarding the capabilities and

adequacy of classical probabilistic safety assessment (PSA) techniques (fault/event tree) to

account for dynamic system interactions and time-dependent accident sequence evolution.

Subsequently there is an interest within the PSA community to complement the classical

techniques with dynamic methodologies; driven in addition by a goal to eventually develop

a framework for integrated safety assessment. The first phase of this research investigates

and addresses the limitations of classical techniques and performs a methodological

comparison between classical and dynamic PSA techniques. The dynamic flowgraph

method (DFM) and Markov model coupled with cell-to-cell mapping technique (Markov-

CCMT) were the two dynamic methodologies selected for this research. These methods are

ranked as the top methodologies with favorable and minimal uncertainties as determined

by the United States Nuclear Regulatory Commission. The capabilities and limitations of

the techniques are demonstrated by applying it to a benchmark liquid level control system

exhibiting dynamic characteristics and interactions. Reliability analysis of the system using

ET/FT are performed using the CAFTA code. DFM model of the benchmark system is

developed using the DYMONDA code and coupling of Markov-CCMT model is performed

using Fortran95 and MATLAB code. Classical techniques were found to overestimate the

predicted top event frequencies by more than one order of magnitude depending on whether

or not dynamic interactions among the units through the state variable is accounted for. The

study shows that DFM focuses more on sequential probabilistic system evolution, whereas

Markov-CCMT emphasizes the exact timing of a failure event. The second phase involves

the development of a novel approach for integrated reliability assessment of passive safety

systems in small modular reactors. A stochastic model of a passive Isolation Condenser

System (ICS) was developed, and its state transition probabilities are computed using finite

element method. The analysis predicts high system reliability, with the ICS most likely to

fail by pressure boundary breach followed by condensate return and venting unit failure.

Key words: Dynamic and classical PSA; DFM; Markov-CCMT model; SMRs; passive

safety systems

Page 4: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

iv

AUTHOR’S DECLARATION

I hereby declare that this thesis consists of original work of which I have authored. This is

a true copy of the thesis, including any required final revisions, as accepted by my

examiners.

I authorize the University of Ontario Institute of Technology to lend this thesis to other

institutions or individuals for the purpose of scholarly research. I further authorize

University of Ontario Institute of Technology to reproduce this thesis by photocopying or

by other means, in total or in part, at the request of other institutions or individuals for the

purpose of scholarly research. I understand that my thesis will be made electronically

available to the public.

Chireuding Zeliang

Page 5: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

v

STATEMENT OF CONTRIBUTIONS

I hereby certify that I am the sole author of this thesis. The current research contributes in

uncovering the limitations and capabilities of classical probabilistic risk assessment

techniques used in nuclear power plants; and demonstrates the advantages of dynamic

probabilistic risk assessment techniques (Chapter 4), and how some risk-significant

scenarios may not be captured by classical techniques. The work described in Chapter 5

presents a novel approach for integrated risk assessment of passive safety systems in small

modular reactor technology.

Page 6: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

vi

ACKNOWLEDGEMENT

I express my sincere gratitude and appreciation to the Dean and Professor Akira Tokuhiro

for his invaluable guidance and continuous support throughout my research activities. His

advice on my research as well as my professional career have been priceless. I would not

have imagined a better advisor and mentor for my research works.

I would like to extend my gratitude to Professor Lixuan Lu for her advice and insightful

contributions to my research pathways.

My sincere acknowledgement goes to the International Atomic Energy Agency and

University of Ontario Institute of Technology for accepting my research proposal, partially

funding my research, and providing me the opportunity to work as the Co-Scientific

Investigator of the IAEA-UOIT Coordinated Research Project.

I am highly indebted to my mentor, Dr. Atam Rao, without whose support and

encouragement, I would not have made it this far. I shall ever remain grateful to you.

I thank all the other members of my advisory committee for their constructive comments

and suggestions.

Last but not the least, words cannot express how grateful I am to my family for the constant

support, encouragement and sacrifices that they have made on my behalf. To my parents,

you have been and continue to be an inspiration in my life.

This thesis is dedicated to my parents and siblings.

Page 7: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

vii

TABLE OF CONTENT

ABSTRACT ...................................................................................................................... III

ACKNOWLEDGEMENT ................................................................................................ VI

TABLE OF CONTENT ................................................................................................... VII

LIST OF FIGURES .......................................................................................................... IX

LIST OF TABLES ............................................................................................................ XI

LIST OF ACRONYMS .................................................................................................. XIV

NOMENCLATURE ........................................................................................................ XV

CHAPTER 1: INTRODUCTION ........................................................................................ 1

1.1. Background ............................................................................................................ 1

1.2. Motivations ............................................................................................................ 8

1.3. Research Objectives ............................................................................................... 9

1.4. Scope of the Research ............................................................................................ 9

1.5. Thesis Organization ............................................................................................... 9

CHAPTER 2: LITERATURE REVIEW ........................................................................... 11

CHAPTER 3: PROBABILISTIC SAFETY ASSESSMENT TECHNIQUES .................. 24

3.1. Fault Tree Analysis .............................................................................................. 24

3.1.1. Qualitative Evaluation of Fault Trees ..................................................... 26

3.1.2. Quantitative Evaluation of Fault Trees ................................................... 28

3.2. Event Tree Analysis (ETA) ................................................................................. 29

3.3. Dynamic Flowgraph Method ............................................................................... 32

3.3.1. Theoretical Basis .................................................................................... 32

3.3.2. Method of Generalized Consensus ......................................................... 35

3.3.3. DFM Modelling Process ......................................................................... 41

3.3.4. Timed-Fault Tree Construction .............................................................. 43

3.4. Classical Markov Chain ....................................................................................... 44

3.4.1. Formulation of the Markov Chain .......................................................... 44

3.4.2. Computation of State Probabilities ......................................................... 45

3.5. Cell-to-Cell Mapping Technique ......................................................................... 48

3.6. Coupled Markov-CCMT Model .......................................................................... 52

CHAPTER 4: APPLICATION OF CLASSICAL AND DYNAMIC METHODOLOGIES

...................................................................................................................... 59

Page 8: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

viii

4.1. The Benchmark System Description ................................................................... 59

4.2. Fault Tree Analysis of the Benchmark System ................................................... 63

4.3. Event Tree Analysis of the Benchmark System .................................................. 66

4.4. Markov Model of the Benchmark System ........................................................... 68

4.4.1. Qualitative Assessment........................................................................... 74

4.4.2. Quantitative Assessment......................................................................... 77

4.5. DFM Model of the Benchmark System ............................................................... 81

4.5.1. DFM Model results ................................................................................. 85

4.5.2. Timed-Fault Tree Generation from DFM ............................................... 90

4.6. Markov-CCMT Model of the Benchmark System .............................................. 94

4.7. Conclusion and Comparison .............................................................................. 110

CHAPTER 5: AN INTEGRATED APPROACH FOR RELIABILITY ASSESSMENT

OF PASSIVE SAFETY SYSTEM ............................................................. 120

5.1. Introduction ........................................................................................................ 120

5.2. Research Project Roadmap ................................................................................ 120

5.3. An Integrated Framework for Dynamic Reliability Assessment of Passive Safety

Systems .............................................................................................................. 125

5.4. The Benchmark Passive Isolation Condenser System ....................................... 132

5.5. The ICS system characterization ....................................................................... 139

5.6. Markov Model of the Benchmark Passive ICS ................................................. 154

CHAPTER 6: SUMMARY AND CONCLUSIONS ....................................................... 167

6.1. Conclusions ........................................................................................................ 167

6.2. Recommendations for Future Work .................................................................. 170

REFERENCES ................................................................................................................ 172

APPENDIX I ................................................................................................................... A-I

APPENDIX II ................................................................................................................. A-V

APPENDIX III ............................................................................................................ A-VIII

APPENDIX IV ............................................................................................................. A-XII

APPENDIX V ............................................................................................................ A-XVI

APPENDIX VI .......................................................................................................... A-XVII

Page 9: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

ix

LIST OF FIGURES

Figure 1: Flowchart of the research approach ..................................................................... 4

Figure 2: Example fault tree (pressure tank) ..................................................................... 25

Figure 3: Example event tree ............................................................................................. 30

Figure 4: Basic building blocks of DFM Model ............................................................... 42

Figure 5: Markov state-transition diagram of a two state component ............................... 46

Figure 6: System reliability and unreliability for a two state non-repairable system........ 47

Figure 7: State-transition diagram of a two state component with repair ......................... 47

Figure 8: State probabilities for component with repair .................................................... 48

Figure 9: A two dimensional state space discretization .................................................... 50

Figure 10: Coupled Markov-CCMT flowchart for integrated system analysis ................. 53

Figure 11: The benchmark liquid level control system ..................................................... 60

Figure 12: Fault tree for the benchmark system failure (binary state) .............................. 63

Figure 13: Fault tree for the benchmark system drained ................................................... 64

Figure 14: Fault Tree for the Benchmark System Overflow ............................................. 65

Figure 15: Event tree for initiating event “Unit-1 Failed-Closed” .................................... 67

Figure 16: Transition diagram from Layer-1 to Layer-2 ................................................... 71

Figure 17: Transition path from Layer-0 to Unit-1 failed-closed to Layer-3 .................... 72

Figure 18: System state transition for Unit-1 failed-open ................................................. 75

Figure 19: DFM model of Total Liquid Flow from Unit-1 ............................................... 83

Figure 20: DFM Model of the Benchmark System ........................................................... 83

Figure 21: Sensitivity of number of PIs to discretized state variable intervals ................. 90

Figure 22: Timed-fault tree top event and its transition from @t=0 to @t= -1 ................. 91

Figure 23: Final timed-fault tree after consistency check ................................................. 93

Figure 24: Proposed coordinated research project roadmap ........................................... 121

Figure 25: Stepwise review and flowchart for design and system selection ................... 123

Figure 26: The project overview and methodology flowchart ........................................ 127

Figure 27: Representation of passive safety system in classical ET/FT ......................... 129

Figure 28: Passive Isolation Condenser system .............................................................. 133

Figure 29: Markov transition diagram of the condensate drain unit ............................... 142

Figure 30: Fault tree of the condensate drain unit ........................................................... 144

Page 10: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

x

Figure 31: Markov transition diagram of the vent unit ................................................... 146

Figure 32: Markov state transition diagram for the HX unit ........................................... 149

Figure 33: Markov state transition diagram of the water makeup unit ........................... 152

Figure A-I-1: Second-order failures state transition diagram ........................................... A-i

Figure A-I-2: Markov state transition diagram for second case ..................................... A-ii

Figure A-I-3: Top event comparison between FT and Markov model .......................... A-iii

Figure A-III-1: Timed-fault tree for transfer gate G5 ..................................................... A-iv

Figure A-III-2: Timed-fault tree for transfer gate G14 .................................................... A-v

Figure A-III-3: Timed-fault tree for transfer gate G15 .................................................... A-v

Figure A-III-4: Timed-fault tree for transfer gate G7 ..................................................... A-vi

Figure A-III-5: Timed-fault tree for transfer gate G8 ..................................................... A-vi

Figure A-III-6: Reduced timed-fault tree after consistency check ............................... A-vii

Page 11: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

xi

LIST OF TABLES

Table 1: A summary of dynamic PSA techniques ............................................................. 15

Table 2: Summary of review of DFM for dynamic PSA ................................................... 18

Table 3: Summary of review of Markov-CCMT model for dynamic PSA ....................... 22

Table 4: Boolean algebra binary reduction rules ............................................................... 26

Table 5: Accident sequence from the event tree ................................................................ 31

Table 6: Possible sensor states ........................................................................................... 36

Table 7: Liquid level in the tank and measured level ........................................................ 36

Table 8: Output decision table- the measured liquid level................................................. 36

Table 9: State variable discretization scheme .................................................................... 60

Table 10: Control laws for the benchmark system ............................................................ 61

Table 11: Total net fluid flow into/out of the tank ............................................................. 62

Table 12: Possible unit state combination ......................................................................... 70

Table 13: Transitional layers representation of system states............................................ 71

Table 14: A qualitative comparison of three different approach ....................................... 77

Table 15: Transition probabilities in terms of Pn(t) .......................................................... 80

Table 16: System hardware identification ......................................................................... 81

Table 17: System parameter identification ........................................................................ 82

Table 18: System conditioning node identification ........................................................... 82

Table 19: Discretization of TL into 5 intervals .................................................................. 84

Table 20: Decision table for TT1 with 5 discretized liquid level intervals........................ 84

Table 21: Transition table after eliminating U2-SB @t=-1 ............................................... 85

Table 22: Quantification of prime implicants for system overflow ................................... 86

Table 23: Complete base for the top event “TL=+2 @t=0” with 2 time steps .................. 87

Table 24: PIs for system overflow “TL=+3 @t= 0” .......................................................... 88

Table 25: Minimal cut-set for the final timed-fault tree .................................................... 92

Table 26: System state transition probabilities .................................................................. 95

Table 27: Possible state transition among distinct unit state combinations ....................... 95

Table 28: Canonical form of the transition probability matric .......................................... 96

Table 29: Fundamental matrix N = (I-Q)-1 ....................................................................... 98

Table 30: Matrix TN .......................................................................................................... 99

Page 12: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

xii

Table 31: Controlled variable discretization scheme ....................................................... 101

Table 32: Cell discretization via equal weight quadrature scheme .................................. 101

Table 33: Pr{g(j/j', n', k∆t)} for k = 1 ............................................................................ 105

Table 34: A small portion of the overall system transition matrix q(n, j/j', n', k∆t) for k =

1 ....................................................................................................................................... 106

Table 35: Top event probabilities for the benchmark system .......................................... 108

Table 36: Sample state transition probabilities ................................................................ 108

Table 37: Predicted failure probability of the BS using FT (binary) ............................... 111

Table 38: Predicted failure probability of the BS using FT (multi-state) ........................ 111

Table 39: Predicted failure probability of the BS using ET ............................................. 111

Table 40: Predicted failure probability of the BS using Markov model with the qualitative

consideration taken in Table 12. ...................................................................................... 112

Table 41: Predicted failure probability of the BS using Markov-CCMT model ............. 112

Table 42: Predicted failure probability of the BS using DFM ......................................... 112

Table 43: Failure modes and rate of the main/bypass valve ............................................ 141

Table 44: Possible drain unit state combination .............................................................. 141

Table 45: Unit state ordering and the inclusion of CCF .................................................. 142

Table 46: Condensate drain unit state transition probabilities ......................................... 143

Table 47: Merged condensate drain unit state ................................................................. 144

Table 48: Failure modes and rate of the vent valve ......................................................... 145

Table 49: Unit state numbering and the inclusion of CCF .............................................. 145

Table 50: Vent unit states transition probabilities ........................................................... 147

Table 51: Failure modes and rate of the heat exchanger.................................................. 147

Table 52: Possible states of the HX unit .......................................................................... 148

Table 53: The overall HX unit state probabilities ............................................................ 150

Table 54: Failure modes and rate of the vent valve ......................................................... 151

Table 55: Unit state combinations and ordering (notations) ............................................ 151

Table 56: Overall makeup water unit states ..................................................................... 153

Table 57: Individual unit notation and numbering........................................................... 154

Table 58: Possible individual unit state combinations ..................................................... 155

Table 59: Systematic organization of system states in Layer fashion ............................. 158

Page 13: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

xiii

Table 60: Transition from Layer 0 to Layer 1 ................................................................. 159

Table 61: Transition from Layer 1 to Layer 2 ................................................................. 159

Table 62: Transition from Layer 2 to Layer 3 ................................................................. 159

Table 63: Transition from Layer 3 to Layer 4 ................................................................. 160

Table 64: Transition from Layer 4 to Layer 5 ................................................................. 160

Table 65: Sample ICS state transition probabilities with 10 time steps .......................... 166

Page 14: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

xiv

LIST OF ACRONYMS

BE : Basic Event

CCF : Common cause failure

CCMT : Cell-to-cell Mapping Technique

CRP : Coordinated Research Project

DFM : Dynamic Flowgraph Method

ET : Event Tree

FMEA : Failure modes and effect analysis

FT : Fault Tree

HX : Heat Exchanger

IAEA : International Atomic Energy Agency

ICS : Isolation Condenser System

IE : Initiating Event

iPWR : Integral Pressurized Water Reactor

MCS : Minimal cut-set

MVL : Multi-valued Logic

NPPs : Nuclear Power Plant

PI : Prime Implicants

PRA : Probabilistic Risk Assessment

PSA : Probabilistic Safety Assessment

PSSs : Passive Safety Systems

RCS : Reactor Coolant System

RPV : Reactor Pressure Vessel

SMRs : Small Modular Reactors

US NRC : United States Nuclear Regulatory Commission

Page 15: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

xv

NOMENCLATURE

l : Liquid level in the tank

��𝑙, ��𝑙 : Lowest and highest allowable magnitudes of state variable l

𝑐𝑚(𝑛𝑚 𝑛′𝑚⁄ , 𝑘∆𝑡) : Transition probability for Unit-𝑚 from state 𝑛′𝑚 → 𝑛𝑚 during the

time step [𝑘∆𝑡, (𝑘 + 1)∆𝑡]

𝑓𝑛 and 𝑓𝑛′ : Flowrate for the component state combination 𝑛 and 𝑛′

𝑓𝑛(𝑥) : L-vectors with components 𝑓𝑛,𝑙(𝑥)

𝑔(𝑗 𝑗′, 𝑛′, 𝑘∆𝑡⁄ ) : Pr{Controlled state variables are in cell 𝑗 at time (𝑘 + 1)∆𝑡 given

that the controlled variables are in cell 𝑗′ at time 𝑡

ℎ(𝑛 𝑛′⁄ , 𝑘∆𝑡) : Conditional probability that the unit state combination at time 𝑡 =(𝑘 + 1)∆𝑡 is at 𝑛, given that the unit state combination at time 𝑡 =𝑘∆𝑡 is at 𝑛′

ℎ(𝑛 𝑛′⁄ , 𝑗′ → 𝑗, 𝑘∆𝑡) : Pr{component state combination is in state 𝑛 at time 𝑡 = (𝑘 +1)∆𝑡 given that the component state combination is in state 𝑛′ at

time 𝑡 = 𝑘∆𝑡, and the controlled variables move from cell 𝑗′ to cell

𝑗 during 𝑘∆𝑡 ≤ 𝑡 ≤ (𝑘 + 1)∆𝑡}

𝐼𝑘 : Number of states for unit k

𝐽 : Total number of 𝑉𝑗

𝐽��, 𝐽𝑟 : Number of 𝑉𝑗 in ��𝑟 and ordering of 𝑉𝑗 in V respectively, 𝐽𝑟 =

𝐽𝑟−1 + 𝐽𝑟 for 𝑟 = 1, 2, … . . . 𝑅; 𝐽𝑜 = 0

L : Number of state variables

M : Number of units in the benchmark system

𝑛 : Component/unit state combination index

𝑛𝑚 : Component/unit state index (𝑛𝑚 = 1, 2, … . . 𝑁𝑚

𝑁𝑚 : Total number of 𝑛𝑚

𝑁 : Number of component state combinations

𝑃(𝑛, 𝑥, 𝑡) : 𝑃𝑟{𝑖(𝑡) = 𝑆𝑛, 𝑥(𝑡) = 𝑥}

𝑃𝑛,𝑗((𝑘 + 1). ∆𝑡) : Probability that the system is in cell 𝑗 and unit state combination

𝑛 at time (𝑘 + 1)∆𝑡

𝑞𝑛,𝑗𝑛′,𝑗′(𝑘∆𝑡) : Elements of the transition matrix for the Markov chain

R : Number of control regions in V

𝑆𝑛 : Ordered set {𝑛𝑖 , … , 𝑛𝑘}

𝑡 : Time

Page 16: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

xvi

V : Control space, 𝑉 ≡ {𝑥;��𝑙 < 𝑥𝑙 < ��𝑙}

�� : Complement of V, (V U ��) = E

𝑉𝑗 : Cells that partition the state variable state space 𝑗 = 1, 2, … . . 𝐽

��𝑟 : Control region r, r = 1, ....., R

𝑉𝛾 : Pairwise disjoint intervals in ��, 𝛾 = 𝐽𝑅 + 1,…… . , 𝐽𝑅 + Γ

𝑣𝑗′ : Volume of the cell 𝑉𝑗′

𝑥𝑙(𝑡) : Magnitude of the state variable l at time t (l= 1, 2, ….., L)

𝑥(𝑡 + ∆𝑡) : State vector magnitude at time (𝑡 + ∆𝑡) or arrival point

��(𝑘+1)∆𝑡(𝑥′, 𝛼𝑛′ , 𝑘∆𝑡) : Location of the system in the state space at time 𝑡 = (𝑘 + 1)∆𝑡,

given that the system location is 𝑥′ at time 𝑡 = 𝑘∆𝑡 and the

component state combination is at 𝑛′

Γ : Number of system failure types

𝜆𝑚𝑓𝑜/𝜆𝑚

𝑓𝑐 : Failure rates of the units with the subscript failed-open and closed

𝛿𝑛′,𝑛 : Kronecker delta

∆𝑡 : Time step

𝜓𝑖(𝑛) : Probability of the system being in cell 𝑖 at 𝑡 = 𝑛

: Logic AND gate

: Logic OR gate

Page 17: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

1

CHAPTER 1: INTRODUCTION

Background

Nuclear energy provides a viable option in contributing towards a reduced carbon world, and to

meet the ever-increasing global energy demand in a safe, reliable and sustainable way. In nuclear

power plants (NPPs), safety assurance is achieved through the application of defense-in-depth

strategy and a robust regulatory framework in the process of design, commissioning, operation

and decommissioning. Since the inception of nuclear power technology, safety analysis has been

performed to demonstrate and understand the safety of an NPP as well as to meet the stringent

regulatory requirements. An overall safety assessment of a NPP includes deterministic safety

analysis (DSA), hazard analysis and probabilistic safety assessment (PSA) [CNSC-REGDOC-

2.5.2 (2014)]. This thesis focusses on the application of PSA techniques to estimate the risk

associated with an NPP. To assist in the process of an integrated risk-informed decision making,

PSA strive to answer some of the fundamental questions possessed by the application of nuclear

technology, including, “what can go wrong?” “how likely is it?” and “what are its consequences?”

[WASH-1400 (1975)]. Thus, a comprehensive PSA considers the likelihood of a failure event,

accident progression, and consequences resulting from the failure event. Here, failure event may

include equipment failure, human error or any event that challenges the plant normal operation.

This approach provides insights into the strengths and weaknesses of the design and operation of

an NPP as well as give confidence in the design alignment with fundamental safety objectives

[CNSC-REGDOC-2.4.2 (2014)]. Due to the specific features of nuclear installation, PSA is

categorized into three (3) levels: Level 1, Level 2 and Level 3. A Level 1 PSA estimates the

frequency of accidents leading to a core damage, i.e., an estimation of core damage frequency

(CDF). Level 2 PSA estimates the frequency of accidents that release radioactivity from an NPP.

The results obtained from Level 1 forms the basis for Level 2 analysis. Finally, Level 3 PSA starts

with the results obtained from Level 2 i.e., the radioactivity release accidents, and estimates the

accident consequences in terms of injury to the public and environmental damage [IAEA-INSAG-

6 (1992)].

This research focus on Level 1 PSA which models the various plant response to an initiating event

(IE) that challenges the normal plant operation. The plant response implies the

activation/operation/failure of several safety and mitigation systems (also called frontline

Page 18: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

2

systems) following an initiating event. These sequence paths (i.e., success or failure of frontline

systems) is called the accident sequence and may include several sequences for an IE which can

result in a safe reactor state or can lead to a reactor core damage. Historically, the modelling of

such accident sequences in nuclear industry is visualized graphically with the event tree (ET)

technique, in which, the frontline systems are represented as top events that are required to

respond to an IE. Accident sequence modelling with ET can be called as a system level analysis.

In order to determine the root cause of a failure event, the frontline systems are modelled

individually with the graphical fault tree (FT) technique. The FT method constitute a component

level analysis. Finally, an integrated risk assessment model of a plant and an estimate of the

frequency of core damage is accomplished by linking the ET and FT model [WASH-1400 (1975),

NUREG/CR-2300 (1983)]. Besides the FT technique, there exist several other methodologies for

individual system analysis such as the discrete space-time Markov model which is a major part

of this research. In this thesis, risk estimation with the fault tree, Markov model and event tree

methodologies will be categorized under the term classical PSA techniques. The classical

approach is widely accepted and recognized as a standard tool for licensing, regulation and safety

analysis within the nuclear industry. Most of the current existing NPPs PSA model are based on

the classical approach (ET/FT) substantiated with sensitivity, importance and uncertainty analysis

to add credit and increase confidence in the model.

While a high level of safety is maintained with a conservative and an integrated safety assessment

process, the nuclear industry nevertheless has experienced major accidents including, Three Mile

Island accident (1979), Chernobyl accident (1986) and the recent Fukushima Daiichi accident

(2011); all three accidents resulting from human error. The nuclear community has been

incorporating the lessons learned from these major accidents and the practical countermeasures

to cope with such an accident, resulting into an evolving rigorous and stringent regulatory/safety

requirements. This has further led the PSA community to re-examine and question the capabilities

of classical PSA techniques to model human error and adequately account for risk-significant

accident sequence occurring from dynamic system behavior and interactions. In the past decades,

numerous concerns have been raised in several reviews regarding the capabilities of classical PSA

techniques, including: [Sui et al. (1989, 1994); Marzio et al. (1998); Aldemir et al. (2007, 2013);

Devooght et al. (1992(a), 1996; Acosta et al. (1991); Marseguerra et al. (1998); NUREG/CR

6901 (2006); Kirschenbaum et al. (2009)]. In brief summary, the classical PSA techniques lack:

Page 19: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

3

1. Treatment of time element [Devooght et al. (1992(a)); Marseguerra et al. (1996); Marzio

et al. (1998); NUREG/CR-6901 (2006)];

2. Dependencies arising from dynamic interactions [Hassan et al. (1990); Belhadj et al.

(1992)];

3. Limited capability in modelling system/components with multi-states [Acosta et al.

(1991)];

4. Treatment of event sequence ordering [Aldemir (1987)];

5. Inadequate treatment of human interactions/error [Acosta et al. (1993)].

The lack of treatment of the above-mentioned factors arising from dynamic interactions means

that the classical PSA approach may not identify or properly quantify potentially risk-significant

sequences and dependencies between the failure events. Furthermore, such interactions may lead

to coupling between stochastic logical events (e.g., valve openings, pump startups) during an

accident scenario, which can significantly impact the predicted system failure probabilities

[Belhadj et al. (1992)]. These limitations of classical PSA techniques entailed the development

of an advanced class of methodologies commonly called dynamic PSA (DPSA) methodologies

or integrated deterministic and probabilistic safety assessment (IDPSA) techniques. Dynamic

PSA methodologies are defined as those which use a time-dependent phenomenological model

of system evolution along with its stochastic behavior to account for possible dependencies

between failure events [Aldemir (2013); NUREG-6901 (2006)]. The use of the term “dynamic”

in this thesis implies an approach that combines explicit modeling of system deterministic

evolution with stochastic modeling. Dynamic methods explicitly model the time element, state

ordering and the mutual interactions among state variables, system components, and operators.

Furthermore, an adequate assessment of the overall risk requires accounting for aleatory

uncertainties arising from the stochastic nature of component failures as well as epistemic

uncertainties arising from limited knowledge of the physical phenomenon relevant to the system.

Dynamic PSA provide an integrated framework to account for both the uncertainties

simultaneously [Zio (2014)]. Thus, treating the deterministic and probabilistic approach in an

integrated fashion, unlike the classical approach where the two are treated separately. Several

dynamic methodologies have been proposed with most of the methodologies in the developing

phase including and have been reviewed in the literatures [Sui (1994); Labeau et al. (2000);

Page 20: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

4

Aldemir (2013); NUREG-6901]. In this thesis, DFM and Markov-CCMT model is selected for

further research, and demonstration of the features, advantages and limitations of the

methodologies. Figure 1 presents a flowchart of the research approach and methodology

selection.

Figure 1: Flowchart of the research approach

Before further detail discussion into the specific methodologies, the five (5) factors mentioned

earlier will be described briefly in context to the classical approach. It should be noted that this

research is bounded within the first four (4) elements. First, the ET/FT methodologies are based

on static logic approach. Neither does it capture the time of occurrence of a failure event nor the

time distribution before an undesired event is reached, i.e., the time available for the subsequent

failure event following an IE. Due to its static nature, it assumes a system failure as soon as a

minimal cut-set (MCS) occurs; this may lead to overestimating the failure probability. Of course,

time is a significant factor in shaping the system dynamics, i.e., the state variables (e.g., reactor

coolant system pressure) are time-dependent. Second, the classical techniques are neither

developed nor intended to model integrated system dynamics and probabilistic accident

evolution. An accident scenario is simply described by a set of success or failure events, where a

risk-significant accident sequence lead to a single core degraded state. The identification and

NPP Safety Assessment

Event/Fault Tree

methods

Markov Model

Classical PSA

Approach

Dynamic PSA

Approach

Probabilistic Safety

Assessment (PSA)

PSA Level 1

Markov-CCMT

Model Dynamic Flowgraph

Method

Page 21: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

5

analysis of these accident sequence can be performed with ease using the well-known logic based

approach and the laws of probability. Thus accounting for dependencies emerging from dynamic

interactions using the classical approach does not arise. This can lead to an inaccurate estimate of

risk when quantifying risk associated with scenarios for which the system dynamic behavior is a

significant factor [Marseguerra et al. (1998)]. Since this research attempts to model and

adequately account for dependencies, a clear distinction between static and dynamic

dependencies will be discussed along with the dependencies covered by classical PSA techniques.

The principle of defense-in-depth design characteristics in nuclear power plants implies that a

risk-significant accident scenario must involve a failure of multiple barriers, which includes

protection, safety and containment systems. Thus, in order to accurately estimate a plant risk, it

is essential to determine the probability of multiple barrier failures. In probability theory, the joint

occurrence of any two failure events A and B is given by: [Vesely et al. (1981)]

𝑃(𝐴, 𝐵) = 𝑃(𝐴). 𝑃(𝐵|𝐴) = 𝑃(𝐵). 𝑃(𝐴|𝐵)

Where, 𝑃(𝐵|𝐴) and 𝑃(𝐴|𝐵) quantifies the dependencies between event A and B. An overly

conservative or optimistic estimate of the dependencies can result in an error of the estimated risk.

Much effort has been put in to adequately identify and quantify these dependencies between

failure events. Acosta et al. (1991) discuss three type of dependencies treated by classical PSA

techniques: (a) Dependencies arising due to common cause failures; (b) Dependencies due to

functional coupling; and (c) Multiple top events sharing the same basic events or shared

equipment (well treated in the fault tree analysis). However, dependencies outside of these

categories arising due to the dynamic interactions are not well treated by the classical techniques

[Sui et al. (1989)]. These three dependencies are briefly described:

1. Direct dynamic dependencies

These types of dependencies arise from direct interaction between two safety systems through

state variables. A failure of one safety system results in a deviation of the state variables, which

in turn, affects the performance of the other safety system. For example, the sequence of events

that occurred at Hatch 1 plant in 1985 involving a safety relief valve (SRV) and vessel integrity

with the reactor coolant system (RCS) temperature as the state variable.

2. Indirect dynamic dependencies

Page 22: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

6

These dependencies are similar to the direct dynamic dependencies except that there is an

intermediary between the demanded safety systems. The event sequence is as follow: failure of

the first safety system causes a deviation in the state variables, which lead to the intermediate

system to affect the operation of the second safety system. For example, this sequence of events

occurred at the Three Mile Island-II accident, 1979; the failure of a pilot-operated relief valve

(first demanded system) to reclose resulted in a high pressurizer level (the state variable) which

led the operator (intermittent) to throttle the high pressure injection (HPI) system (second

demanded system). The state of HPI is affected by the operator action, which was influenced by

the pressurizer level. The classical approach does not have the feature to account for pressurizer

level-operator interaction.

3. Cyclical dynamic dependencies

These types of dependencies involve the multiple occurrence of certain events. For example,

multiple opening and closing of safety relief valves at Davis-Besse plant, 1985 with the RCS

pressure as the state variable. The affected functions in these sequence of events was the demand

for safety relief valves to open and subsequent demand to reclose due to the variation in RCS

pressure (the state variable). Standard event tree typically models these event sequence (opening

and reclosing) just once in the tree.

Third, the classical approach is based on binary logic (normal or failure), and there exist no well-

defined approach to modelling system/components having multiple states or top events and may

not yield satisfactory result [Hassan et al. (1990); NUREG/CR-6942 (2007)]. For example,

modelling a control valve having the states; normal, fail-closed, fail-open, fail in 50% position,

fail in 80% position, etc. Fault tree tend to model the failure modes independently, rather than

taking into account the competing failure modes. This is especially true while computing the

reliability of a system having several components with multiple failure modes. This phenomenon

is well explained and demonstrated using the FT and Markov model in Chapter 3, Section 3 with

a simple redundant system with two (2) components having three failure modes. We illustrate

how FT approach overestimates the failure probability.

Fourth, in the classical approach, the order in which the frontline system operates following an

IE is normally predefined by the analyst. In the event tree method, the sequencing of failure events

is fixed. For instance, given a demanded system that operates successfully, the sequence is

Page 23: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

7

terminated, i.e., leading to a successful reactor state. However, event sequence can be generated

wherein the system is demanded multiple times due to the system dynamics. Additionally, the

order of occurrence of failure events can lead to a different undesired system state. This is

demonstrated in Chapter 4 where the order of unit failure leads to a system overflow or drained.

As for the fault tree technique, it is not intended to model ordering of failure event, rather it

represents the top event in terms of combinations of basic events (minimal cut-sets) without any

particular ordering index. In contrast, the sequencing of events is not predetermined in dynamic

methodologies but rather is derived from the time-dependent system model solution (system

code) as the system evolves.

Dynamic PSA methodologies thus have features and capabilities to overcome the limitations

encountered in classical PSA techniques. This is especially pertinent when modelling passive

systems where epistemic uncertainties are significant due to lack of knowledge of physical

phenomenon and operating experience, as compared to active systems which have been employed

in most current operating NPPs and have a vast operating experience. This may be especially true

for SMRs with either different forced or free convective systems. The IAEA defines passive

safety systems as “a system that is composed entirely of passive components and structures or a

system, which uses active components in a very limited way to initiate subsequent passive

operation”. [IAEA-TECDOC-626]

Passive safety systems (PSSs) have been employed in several (Gen-III and Gen-III+) reactor

designs [AP1000, ESBWR, NuScale, mPower, SMART] to meet the increasing safety

requirements, to take advantage of natural forces (e.g., gravity), less dependence on active

systems/components (e.g., emergency diesel generators), design simplicity, increased reliability

and economic competitiveness. However, due to their reliance on a small driving force, lack of

data and significant uncertainties; passive safety systems reliability, model development and

performance under normal and accident conditions still remains an open issue [Burgazzi (2009)].

Also, the implementation of passive systems in NPPs is challenging the state-of-the-art safety

analyses (classical PSA/DSA) due to the additional uncertainties rendering a difficult a priori

judgement of the conservative scenarios selected for DSA and PSA [Zio (2014)]. These elements

necessitate for the development and demonstration of consistent methodologies and approaches

for evaluating the reliability of passive safety systems. [Burgazzi (2007, 2009, 2011, 2017);

Marques et al. (2015) Nayak et al. (2014); Zio et al. (2009); IAEA-TECDOC-1752 (2014)]

Page 24: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

8

Motivations

There still exist no consensus on a standardized dynamic PSA technique, unlike the classical

techniques (FT/ET), which are universally accepted all around the world. Rather, most of the

dynamic PSA methodologies are still in the developing and validating stage [Labeau et al.

(2000)]. Furthermore, there is no agreement on the specific application and under what scenario

dynamic PSA methodologies might be appropriate. The motivations for this research thus include:

1. Limitations in classical PSA techniques to account for time element, system interactions,

dynamic dependencies, multistate modelling and state ordering;

2. Dynamic PSA techniques provide an integrated framework by unifying the system

stochastic model as well as its dynamic evolution; thus are anticipated to overcome the

limitations encountered in classical PSA techniques. This is expected and envisioned

within the PSA community to contribute significantly in providing a realistic accident

progression scenario by capturing risk important interactions;

3. Currently there exist no standardized or universally accepted dynamic methodology as

well as a lack of consensus within the PSA community on the applicability and the context

to which dynamic methodologies are required;

4. NUREG/CR-6901 identified and ranked DFM and Markov-CCMT as the top two dynamic

methodologies with the most positive features and least negative or uncertain features;

5. NUREG/CR-6942 recommended thorough research to resolve the challenges for practical

implementation of dynamics PSA methods. The report further went on to state that

“resolving these challenges would involve a stand-alone reliability modeling of full

benchmark system using DFM, Markov/CCMT and classical ET/FT approach”;

6. Dynamic PSA approach has not been applied for integrated reliability assessment of

passive safety systems (to the author’s knowledge). The author was thus motivated to

evaluate the possibility of modelling PSSs using dynamic methods;

7. The supervisor’s experience at NuScale Power, and a full dynamic PRA using system

code (RELAP5) coupling experience for light water reactors (LWRs) under a benchmark

station blackout.

Page 25: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

9

Research Objectives

The objectives of the research are:

1. Formulate the coupling of the Markov model with Cell-to-cell Mapping Technique for

dynamic probabilistic safety assessment;

2. Demonstrate how DFM and Markov-CCMT model can be implemented for modelling

stochastic dynamic system evolution and interactions. This can contribute in performing

system reliability analysis using the two dynamic methodologies as well as uncover the

advantages and disadvantages of the techniques;

3. Illustration of the applicability of backtracking algorithm for timed-fault trees generation

from DFM model;

4. Methodological comparative assessment (qualitative and quantitative) of classical and

dynamic PSA techniques;

5. Development of a novel approach for integrated functional reliability assessment of

passive safety systems in integrated small modular reactors (SMRs).

Scope of the Research

This research consists of two independent areas that will cover the applicability of DFM and

Markov-CCMT model for dynamic probabilistic safety assessment, and development of a novel

approach for integrated functional reliability assessment of PSSs in integrated small modular

reactors. First, the need for dynamic PSA techniques was explored, and two dynamic PSA

methodologies, namely, DFM and Markov-CCMT model was selected for further research and

demonstration purpose. Second, the Markov-CCMT model was selected for reliability modelling

and performance analysis of a passive isolation condenser system (ICS) implemented in a generic

integral pressurized water reactor (iPWR) type SMRs.

Thesis Organization

This thesis consists of six (6) chapters, and is organized as follow:

Chapter 1: This chapter presents the background of the research, motivations for selecting

the research topic, objectives of the research and the structure of the thesis.

Page 26: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

10

Chapter 2: This chapter presents the literature review performed during the course of this

research. The review includes classical PSA techniques, their limitations; thus

the need of dynamic PSA methodologies. A more extensive review of DFM and

Markov-CCMT model is presented.

Chapter 3: This chapter presents the methodologies implemented in this research in a

sequential fashion, which includes, FT/ET analysis, classical Markov model,

DFM and coupled Markov-CCMT model.

Chapter 4: This chapter demonstrate application of methods to a simple liquid level control

system. First, reliability analysis of the benchmark system was performed using

ET/FT and time-dependent Markov model. These form the classical techniques.

Then, DFM and Markov-CCMT model are applied for dynamic reliability

analysis of the benchmark system using DYMONDA and Fortran95 software

and programming tools.

Chapter 5: This chapter presents a novel approach for dynamic reliability assessment of

PSSs. A brief description of the ICS to be used as a benchmark system for the

project is provided. A stochastic model of the ICS and computation of transition

state probabilities is presented.

Chapter 6: Finally, this chapter summarizes the research contributions and also suggests

future research tasks.

Appendix-I: Multi-state modelling with FT and Markov model

Appendix-II: Fortran95 code for the benchmark liquid level control system

Appendix-III: Benchmark system control space discretization in MATLAB using CCMT

Appendix-IV: Benchmark system state vector trajectories results from MATLAB

Appendix-V: Fortran95 code for the passive isolation condenser system

Appendix-VI: Powerpoint presentation used in thesis examination

Page 27: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

11

CHAPTER 2: LITERATURE REVIEW

This chapter presents a literature review that was performed during the course of this research.

Literature review starts with the well-known classical PSA methodologies, their application in

the nuclear industry and available algorithms for minimal cut-set (MCS) computation. A review

of literatures to address limitations of classical methods and currently available dynamic PSA

methods was performed, of which, DFM and the Markov model coupled with CCMT are selected

in this thesis for further research and finally application/demonstration. The literature review is

provided in a sequential order from classical PSA techniques, limitations in classical techniques

and thus, the need for dynamic PSA techniques, and finally a comprehensive review of the

selected methodologies is provided. Literature review during the course of this research was

focused on dynamic methods. Note that in this thesis, usage of the term dynamic PSA and

Integrated Deterministic and Probabilistic Safety Assessment (IDPSA) is synonymous.

A. Classical Probabilistic Safety Assessment Techniques

The classical PSA techniques discussed in this thesis mainly consist of the well-known and

commonly used risk assessment techniques in current NPPs, i.e., the fault tree and event tree

(FT/ET) analysis. The application of classical PSA techniques in NPPs was introduced by the

WASH-1400 (1975) “The Rasmussen Report”, with the objective to identify possible accident

scenarios following an initiating event, and to quantify the likelihood of a reactor core damage

and its consequences by observing the response (success/failure) of the frontline mitigating and

safety systems. WASH-1400 demonstrated the usefulness and practicality of classical techniques.

The risk assessment review group report to the USNRC through NUREG/CR-0400 summarized

that:

“The fault-tree/event-tree methodology is sound, and both can and should be more widely used

by NRC. Proper application of the methodology can therefore provide a tool for the NRC to

make the licensing and regulatory process more rational, in more properly matching its

resources to the risks provided by the proper application of the methodology”.

A comprehensive and systematic approach to the construction, qualitative and quantitative

analysis of fault trees is provided in “Fault tree handbook” by Vesely et al. (1981) and

Page 28: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

12

Stamatelatos et al. (2002). The authors described FTA as an analytical technique whereby an

undesired event or top event is pre-defined by an analyst, and system analysis is performed

backward to uncover the root cause of the top event. The methodology basically consists of three

steps: (1) construction of the tree; (2) qualitative analysis of the tree; and (3) quantification of

minimal cut-sets. A detail guideline to performing a PSA for NPPs via classical techniques is

provided in NUREG/CR-2300 (1983). Fault tree construction (Fussell, 1973) can be done

automatically, for instance, Salem et al. (1976, 1977) developed a new methodology based on

decision tables and implemented in the code Computer Automated Tree (CAT) for automatic

construction of FTs. The analysis of a FT is performed using Boolean algebra, and the top event

is entirely described by an equivalent set of Boolean equation. There exist several algorithms for

computing and quantifying the MCS for a top event, such as the method of obtaining cut sets

(MOCUS) and minimal cut-sets upward (MICSUP) [Rasmussen (1978); Kumamoto et al. (2000);

Ruijters et al. (2015)]. Furthermore, there are several advanced computer codes for automated

construction of FT/ETs, and subsequent generation of MCS/accident sequence including, CAFTA

and FaulTree+ code.

B. Dynamic Probabilistic Safety Assessment Techniques

While classical PSA techniques are accepted as standard safety analysis technique in the PSA

community, numerous concerns have been raised in the recent years regarding its capability to

adequately account for the time element, multistate modelling, state ordering and stochastic

system interaction that shapes the dynamic probabilistic evolution of an accident sequence [Sui

et al. (1989, 1994); Marzio et al. (1998); Aldemir et al. (2007, 2013); Zio (2014); Devooght and

Smidts (1992(a)); Acosta et al. (1991, 1993); Marseguerra et al. (1996)]. The behavior of the

plant state variables resulting in a coupling of failure events, with the failure events being

probabilistically dependent, in contrast to the logical dependencies caused by direct systems

interaction have been debated within the PSA community for several decades [Aldemir et al.

(1987); Sui et al. (1989); Devooght and Smidts (1992(a)); Belhadj et al. (1992); Labeau et al.

(2000)].

Sui (1989) characterized these dependencies into three classes by reviewing the licensee events

reports (1969-1979 and 1985) for incidents that occurred at the US operating commercial light

water reactors, namely: (1) Direct physical dependencies; (2) Indirect physical dependencies; and

Page 29: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

13

(3) Cyclical physical dependencies. Sui underlined the importance of dynamic dependencies and

how it played an important role in the Three Mile Island (TMI-2) accident. Classical PSA

techniques lack the ability to capture these dependencies, and hence may fail to identify risk-

significant scenarios. Four (4) potential dynamic methodologies and their attributes were

discussed for incorporating dynamic behavior of the plant into the accident sequence analysis,

which includes, Expanded Event Tree, Logic Flowgraph Method, Markov model and Dynamic

Logical Analytical Methodology (DYLAM) (Sui et al. (1989)). Sui (1994) further attempted to

address the limitations of classical techniques in accident sequence modelling and suggested

possible dynamic methodologies for explicit treatment of time dependency.

The need for dynamic PSA for accident sequence modelling to account for evolution of the state

variable, operator state of mind and scenario history was further reinforced by Acosta et al.

(1991). The author showed how an overly conservative or optimistic assumptions of failure events

could result in an inaccurate risk estimate. The coupling among the mentioned factors was

demonstrated by applying the Dynamic Event Tree Analysis Method to a SG tube rupture accident

and concluded that the dynamic method can uncover risk-significant scenarios and better define

dependencies between failure events as compared to the classical methods, Acosta et al. (1993).

Marzio et al. (1998) discussed the specific field of application of the dynamic PSA methods, and

the viability of the Monte Carlo technique as a tool to model the stochastic part of the analysis to

reduce computation time. The importance of dynamic methods in PSA modelling of the passive

safety system was underscored and demonstrated with two simple examples (Aldemir (2013)).

On the other hand, Devooght and Smidts (1992(a)) took an analytical approach and formalized

the concept of probabilistic reactor dynamics by providing a rigorous mathematical framework

of continuous event tree in which the deterministic aspect of the reactor state variable trajectories

in state-space is supplemented by the stochastic nature of the reactor configuration (e.g., random

component failure, human error, etc.). The system dynamics is described by a set of partial

differential equations from the Chapman-Kolmogorov equation assuming a Markovian system.

Smidts and Devooght (1992) demonstrated the methodology by applying it to a realistic accident

transient in a fast reactor primary coolant system. The capability of the method to model complex

interaction between operators and the reactor during a transient is present in [Devooght and Smidts

(1992(b)); Smidts (1992)]. However, the applicability of the method to realistic systems is limited

Page 30: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

14

due to its computational demand and continuous time-dependent plant data requirements. A

review of the achievements of the theory of probabilistic dynamics and its adjoint formulation to

determine the outcome of a transient is presented in Devooght and Smidts (1996).

A comprehensive review of the state of dynamic PSA methodologies for reliability analysis of

digital systems in NPPs is reported in NUREG/CR- 6901 (2006). DFM and dynamic event tree or

Markov model approach was ranked as the top two methodologies with the most positive features

and least negative or uncertain features, meeting most of the requirements, and with each

methodology having some advantages as well as limitations. For a proof of concept, a benchmark

digital feedwater control system is analyzed using DFM and Markov-CCMT in the reports

NUREG/CR-6942 (2006) and NUREG/CR-6985 (2009) as a follow-up of NUREG/CR-6901

(2006). A detail review of currently available dynamic PSA techniques, and a modular approach

to the development of high-level and user-friendly tools to implement dynamic methods in

industries is proposed by Labeau et al. (2000). The literature underlined some important issues,

which includes, a significant amount of research is required for the development of the

computational algorithm and the determination of an optimal computational engine. Zio (2014)

laid out the concept, challenges and research directions for industrial application of IDPSA. The

author also noted that, IDPSA should be consider as a complement to the existing DSA and PSA

and not as an alternative. Zio (2014) and Aldemir (2013) showed how IDPSA can explore a more

complete scenario space and coverage of undesired events by integrating a system dynamic and

stochastic aspect, and consistent treatment of uncertainties (aleatory and epistemic) in the

analysis. Table 1 groups and summarizes dynamic PSA techniques into three (3) categories,

namely, (1) Extension of classical methodologies; (2) Explicit state-transition methods; and (3)

Implicit state-transition methods.

Page 31: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

15

Table 1: A summary of dynamic PSA techniques

Parameters

Dynamic PSA Approaches

Extension of classical

PSA methodologies

Explicit state-transition

methods

Implicit state-transition

methods

Input

requirements

All dynamic techniques require: (1) Time-dependent system model; (2) System

configurations under normal/abnormal operating scenario; and (3) Transition

probabilities among these system states

Methodologies Dynamic Flowgraph

Method; Digraph-based

fault tree; Expanded

Event Tree; GO-FLOW

Explicit Markov chain

models; Markov-CCMT

model; Event Sequence

Diagrams

Continuous Event Trees;

Dynamic Event Trees;

Discrete event (Monte

Carlo) simulation

Technique An extension of classical

approach incorporating

system dynamics

Discretized version of

Continuous Event Trees

Analytical based

Accounting for

system

dynamics

Limited capability Explicitly account for

system dynamics and

control laws

Explicitly account for

system dynamics and

control laws

State ordering Sequential top event

ordering in tree structure

A priori explicitly

ordered system states

Evolving sequence from

probabilistic system

dynamics

Scenario history Explicit representation

of scenario history in the

tree structure

Markovian approach-

memoryless

Capability for treatment

of scenario history

Recent

development to

address some

drawbacks

Reduction technique for

component combinatory

explosion

State merging technique

employed to counter

state space explosion

Truncation rule by

probability or event

type; intelligent

sampling scheme to

reduce run time

Treatment of

time element

Limited treatment of

time dependencies

Explicit treatment of

time element

Explicit treatment of

time element

Page 32: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

16

Complex system

modeling

Limited Limited Enables treatment of

complex/realistic models

Integrated user

friendly codes

Several integrated

platform exist, e.g.,

Dymonda, GO-GLOW.

None. Requires code

coupling

None. Requires code

coupling

Operator

behavior model

Limited capability to

model state variables

and operator behavior

Limited capability to

model operator states

Explicitly deals with

operator states and

dynamic man-machine

interface systems

Post-processing Output similar to that of

classical techniques

Requires considerable

post processing

Requires considerable

post processing

Advantages Phase mission scenario

implementation

Ease of treatment of rare

events

Eliminates expert

judgement

Drawbacks No direct treatment and

computation of common

cause failure and

importance measure

Requires explicit

evaluation of transition

probability matrix;

Difficult to envision the

set of all possible system

states prior to scenario

development

Requires significant

computational time and

resources to run realistic

models; Requires

algorithm development

specific to system under

analysis (e.g., situation

specific rules for

operator state transition)

Note: All dynamic PSA methods encounter the phenomenon of state space explosion, however

manifested in different ways for different approaches

B.1. The Dynamic Flowgraph Method

The DFM is a multi-valued logic (MVL) diagraph-based method that express the logical and

dynamic behavior of a system in terms of causal relationships among system parameters, and a

series of discrete state transitions [Garrett et al. (1995); Yau et al. (1995, 1998); Guarro et al.

(1996)]. DFM is a three-step approach: model development, model analysis (qualitative) and

Page 33: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

17

finally the quantitative analysis which includes uncertainty, sensitivity analysis, etc. All the above

three steps can be performed in a single platform using the code DYMONDA (ASCA Inc., 2013).

Once the model is developed, analysis or generation of prime implicants (PIs) can be performed

via an inductive or a deductive algorithm. Modelling of a system involves the construction of

decision tables, which is an extension of the well-established truth table. Application of decision

tables to FT construction was introduced by Salem et al. (1979), and the application of MVL

decision table for risk analysis was comprehensively discussed in Ogunbiyi (1981b). MVL rules

implemented for analysis of critical transition table to obtain the complete base is discussed in the

literatures Ogunbiyi et al. (1981(a), 1981(b)) and Yau (1997). The generation of PIs complete

base can be performed using several methodologies, which includes [Garribba et al. (1985)]: (1)

Tabular method; (2) Nelson method; and (3) Method of Generalized Consensus. Since the

DYMONDA code is based on the Generalized Consensus method, the same methodology is used

for the current thesis. Identification of failure events using PIs and the need for generation of

complete base of PIs are presented in Yau (1997).

A comprehensive and integrated framework for modelling and analysis of system reliability and

safety assurance is captured in the literature [Garrett et al. (1995, 2002); Yau et al. (1995, 1997);

Guarro et al. (1996)]. The integrated DFM approach was first applied for reliability assessment

and verification purpose of a software-driven embedded system by Garrett et al. (1995). Early

on, the methodology was mainly aimed at identifying an optimal testing strategy based on system

behavior analysis, Garrett et al. (1995). A further demonstration of the methodology was

performed using the Titan II space launch vehicle digital flight control system with the objective

to complement classical PSA techniques, Yau et al. (1993). The authors showed how time-

dependent behavior and switching logic can be captured, and how certain postulated events

(desired/undesired) that can occur in a system may be identified by the DFM model. Yau et al.

(1998) then applied the methodology to a realistic PWR steam generator level control system, in

which the author used a fault injection technique to check as to whether the failure event can be

detected. The USNRC developed a full-scale DFM software tool to model, analyze, and test

software design to provide a high level of safety assurance for software-based control systems,

the details of which can be found in NUREG/CR-6465. This NRC report presented distinct

features of the methodology and demonstrated by applying it to a generic process system as well

as to a full scale SG level control system implemented in current NPPs. NUREG/CR-6942 report

Page 34: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

18

(as a follow-up of NUREG/CR-6901) further applied the methodology to model a digital

feedwater control system and to check the extent to which the methodology meets the

requirements set-out in NUREG/CR-6901. The report also provided a framework on how PIs from

DFM can be integrated into an existing PSA model. Finally, generation of timed-FTs from a DFM

model is discussed in works by, [Guarro et al. (1996); Yau (1997); Zeliang et al. (2017)].

Table 2: Summary of review of DFM for dynamic PSA

Authors Contributions Limitations

Salem et al., 1979 Computerized automatic

construction of fault tree using

decision table

Lacks the treatment of multistate

variables; requires analyst

judgement and significant

amount of time for construction

of component decision tables

Ogunbiyi et al.

[1981(a); 1981(b)]

Extended the binary Boolean

algebra to encompass multistate

variables, and simplify consensus

algorithm for non-coherent

systems

Only a theoretical framework;

Applicability and efficiency of

the approach to treat realistic

systems remains an open issue

Garribba et al., 1985 Provided a theoretical framework

for treatment of critical transition

table using Tabular method,

Nelson algorithm and generalized

consensus theory

Lacks the discussion as to which

systems these theories are

suitable and applicable.

Yau et al. 1993,

1995

Demonstrated the capabilities of

DFM by applying it to a realistic

Titan II space launch vehicle

digital flight control system

Focused mostly on the

construction of decision tables

for software subroutines using

Newton-Raphson method

Garrett et al. (1995,

2002)

Provided an integrated modelling

framework for reliability

assessment and verification of

embedded control systems

Focused only on risk assessment

of software driven systems and

its testing strategy

Page 35: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

19

Guarro et al. (1996);

Yau (1997);

NUREG/CR-6942

Provided a theoretical and

systematic framework for

generation of timed-fault tree

from DFM model

Limited discussion on specific

application of dynamic and

physical consistency rule, and

treatment of logic loops

Yau et al. (1997,

1998)

Applied the methodology to a

realistic PWR software driven

SG level control system

Requires further inductive

analysis to locate fault condition,

i.e., backtracking must be

employed even after obtaining

the complete base

USNRC-

NUREG/CR-6465

Developed a full-scale DFM tool

for safety analysis of control

software in advanced reactors

Focused mainly on assurance and

design verification of software

driven closed loop systems,

rather than identification of

failure events

USNRC-

NUREG/CR-6942

Dynamic reliability modelling of

benchmark digital feedwater

control system

Only a proof of concept;

acceptability of failure data;

requires further research to

validate the practicality of DFM

Zeliang et al. (2017) Generated timed-fault tree from

DFM model of a PWR SG level

control system

The approach requires significant

amount of time for systematic

post processing of critical

decision tables; particularly

inefficient in the treatment of

logic loops, where an event can

be backtracked to itself, and

consistency rules must be

checked in every steps.

B.2. Coupled Markov/Cell-to-cell Mapping Technique

Markov Chain: In 1907, A. A. Markov developed the Markov chain or the discrete Markov

process that consisted of a simple chain with infinite sequence of random variables connected in

Page 36: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

20

such a way that 𝑥𝑖+1 for any 𝑖 is independent of the past evolution if 𝑥𝑖 is known (Basharin et al.

(2004)). The Markov chain is a well-established method and can be found in many standard

references [Pukite (1998); Privault et al. (2013); Taylor et al. (2013), chapter-8, 9 and 10;

Sheskin (2016)]. The modelling of dependencies and common cause failures between failures of

component in redundant systems using a continuous time, four (4) state Markov chain was

discussed and illustrated by Platz (1984). Practical application of the methodology to real systems

was unrealized due to a very large transition matrix, or the phenomenon called state space

explosion. Papazoglou (1977) laid down two novel approaches to deduce large transition

probability matrices by systematic state ordering and merging of processes for system exhibiting

symmetries respectively. Several algorithms were developed for systematic generation and

treatment of structural properties of system transition matrix in a computer aided environment

[Amoia et al. (1981); Cafaro et al. (1986); Lesanovsky (1988)].

Cell-to-cell Mapping Technique: The theory of cell-to-cell mapping was first formulated by Hsu

(1980a, 1980b) to describe global behavior of non-linear dynamic systems, in which, a state

variable evolution is treated as the probability of transition among a collection of computational

cells in the state-space. The theory was based on the idea of describing a dynamic system

governed by ordinary differential equations (ODEs) with a finite number of computational cells.

Hsu (1981) generalized the theory of cell-to-cell mapping to treat complex non-linear global

behavior pattern by enabling a cell to have multiple transition probabilities or image cells. The

generalize cell mapping was deduced to the well-known Markov chain in which, the dynamical

properties of the system under analysis is entirely constituted in the transition probability matrix.

The application and improvements of cell mapping technique for determination of global long-

term behavior of non-linear dynamic systems using several approaches such as, cell refinement

and lumping can be found in [Mo-Hong (1993); Spek (1994); Chen (2004)].

Coupling Markov model with CCMT: Markov-CCMT model describes a system dynamics as

probabilistic evolution of state variables in discrete space-time conditioned on system

configuration, Aldemir (1987). The multi-state system components and configuration is modelled

by a Markov chain, and CCMT describes the system dynamics in terms of probability of transition

among a finite set of computational cells in discretized state space and time [Hassan et al. (1990);

Page 37: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

21

Belhadj et al. (1992); NUREG/CR-6942 (2007)]. A comprehensive theoretical basis of the

methodology can be found in [Aldemir (1987); Dinca (1997)].

The concept of coupling Markov model with CCMT was first formulated by Aldemir (1987) to

describe the probabilistic evolution of system behavior in discrete space-time. The process of

coupling with sets of ODEs and its simulation by a Markov chain was demonstrated by applying

it to a hypothetical process control system failure. Aldemir (1989) further illustrated the

limitations of static methods in failure modelling of process control systems and how the overall

failure probability may be overestimated. Hassan et al. (1990) applied the methodology to a

realistic closed loop control system i.e., the high-pressure injection system and SRVs in BWRs

and observed the probabilistic behavior of the systems following a small loss of coolant accident

(SB-LOCA). The authors demonstrated that cell-to-cell transition probabilities can be computed,

provided a data base of system behavior under normal and abnormal operation. Belhadj et al.

(1992) further demonstrated capabilities of the methodology in modelling time-dependent

behavior by applying it to a feed-bleed cooling scenario in BWRs through intermittent operation

of high pressure core spray system and SRVs following a SBLOCA. In an effort to reduce the

computational time and computer storage requirements, Belhadj et al. (1995) developed two

algorithms based on the Chapman-Kolmogorov equation. Accounting for uncertainties in system

parameters and initial conditions was demonstrated by Aldemir et al. (1996). Tombuyses et al.

(1997) performed a parametric study on computational efficiency of continuous CCMT and

concluded that a fourth order Runge-Kutta and mid-point implicit scheme seems to be suitable

for short and steady state system behavior simulation.

A benchmark system analysis i.e., a typical SG feedwater control system in PWR to be used as a

basis of the methodology was proposed in NUREG/CR-6942 (2007) and Kirschenbaum et al.

(2009). Gomes et al. (2013) demonstrated the superiority of the methodology in capturing

interactions and identifying possible failure events in digital systems by implementing it to a

typical PWR digital SG level control system. Yang et al. (2016) developed an algorithm for

deductive implementation of the model, and concluded that, the approach in principle is similar

to FT, in that, it starts with a predefined top event and backtrack for root cause analysis.

Page 38: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

22

Table 3: Summary of review of Markov-CCMT model for dynamic PSA

Authors Contributions Limitations

Aldemir (1987,

1989)

Coupled classical Markov model

with CCMT for dynamic

reliability assessment and failure

modelling of process control

system

Computationally intensive and

impractical with increasing

failure modes and state variables;

only for binary failure modes

Aldemir et al (1996) Theoretical framework for

treatment of uncertainties in

system parameters and initial

conditions

Considerable resource

requirements, and a very fine

partitioning of state space

required for uncertainty analysis

Hassan et al. (1990) Observed that failure

characteristic maybe dependent

on the order of failure and exact

timing of failure event

The approach is useful only if

there is an existing data base of

state variable evolution

Belhadj et al. (1992) Demonstrated that the predicted

failure probability may

significantly change depending

on whether dynamic interactions

is accounted for or not.

Classical approach (FT) is used

for state merging purpose,

therefore not completely dynamic

in nature

Tombuyses et al.

(1997)

Computational efficiency of

CCCMT maybe increased using

4th order Runge-Kutta and mid-

point implicit scheme for short

and long term system behavior

respectively

Predicted results are not validated

or compared with CCMT space-

time discretization schemes;

approach not suitable for

increased step size as noted by

the authors.

Kirschenbaum et al.

(2009)

Applied the methodology to a

realistic feedwater control system

for operating PWR

Quantification of top event

probability was not performed

Page 39: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

23

Gomes et al. (2013) Applied the methodology to a

realistic SG level control system

in a typical PWR

The study only focused on the

main and backup computers, and

not on mechanical components

(e.g., control valves and pumps)

Yang et al. (2016) Proposed a theoretical framework

for deductive implementation of

the methodology

The approach is applicable only

when the transition probability

matrix is symmetric

Page 40: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

24

CHAPTER 3: PROBABILISTIC SAFETY ASSESSMENT TECHNIQUES

In this thesis, several PSA methodologies was studied, and are categorized into two parts-

Classical PSA and Dynamic PSA methodologies. The classical methods consist of the well-

established static ET/FT analysis, whereas dynamic methods consist of DFM and Markov-CCMT

model. More discussion is provided on dynamic PSA methodologies, and a detail theoretical basis

of Markov-CCMT model was emphasized due to its nature of tight coupling between stochastic

and dynamic model. Several examples are performed to get an intuitive understanding of the

methodologies as well as to make a direct comparison among the methodologies. Sub-section 3.1

and 3.2 discusses the FT/ET technique respectively; Sub-section 3.3 presents the DFM technique;

Sub-section 3.4 discusses the classical Markov chain; Sub-section 3.5 provide a theoretical basis

for CCMT, and Sub-section 3.6 provide a detail discussion of Markov-CCMT model.

Fault Tree Analysis

Fault tree analysis (FTA) is a well-established logic method, and one of the most prominent

probabilistic technique used in probabilistic safety assessment and reliability assessment of

engineered systems in NPPs. FT is a logical, systematic and comprehensive approach for system

safety analysis that is capable of uncovering design and operational weakness [Stamatelatos et

al. (2002)]. The methodology starts by defining an undesired system state (top event) and is traced

back to the basic events to determine the possible causes of the predefined top event. A basic

event (BE) which normally is a failure event can be a failure of system components, human error,

software error, or any other events which can lead to the top event. A FT thus depicts the logical

interrelationships of BEs that lead to the top event of the FT, i.e., it models the propagation of

failures through a system, and how these failure interactions can lead to a total system-failure

mode. Thus, for a given FT, the top event is characterized by a set of BEs, in that, if all BE occurs,

the top event is expected to occur. These set of basic events is called the “cut-set”. A cut-set can

be further deduced to a “minimal cut-set (MCS)”, such that if any BE is removed from the set, the

remaining BEs collectively are no longer a cut-set. There exist several approach for determination

of MCS based on Boolean algebra, and the process of deduction is called the qualitative analysis.

Thus, the objective of the methodology, or an extensive effort is put in to determine the MCSs

and its corresponding probabilities that ultimately dictates the occurrence probability of a top

Page 41: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

25

event. Propagation of MCSs given the BE failure probabilities through the tree for computation

of top event probability is called quantitative analysis. It should be noted that the fundamental

concept of the methodology is based on binary logic, i.e., a component can either be in normal or

failed state only. Hence, the top event can only take two states since it is a function of BEs.

For illustration and comprehension purpose, a generic FT taken from Vesely et al. (1981) and

constructed in CAFTA [EPRI, 2013] is shown in Figure 2. The system consists of a pressure

tank, pump and associated control system that regulates the tank pressure by controlling the pump.

When the pressure setpoint is reached in the tank, the control system opens the pressure switch

and thus removing power to the pump, causing the pump to stop operation. A failure of the

pressure switch or relay contacts activates an alarm that enables the operator to manually switch-

off the pump. Of course, a combination of these failures will guarantee a system failure (tank over

pressurized) as shown in the FT.

Figure 2: Example fault tree (pressure tank)

Page 42: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

26

The FT shown in Figure 2 is compose of the top event, intermediate events (1 to 4), logic gates

(G1 to G4 including the TOP gate) and basic/primary events (BE1 to BE6). Firstly, the BEs are

those that requires no further development and are represented by a circle in the fault tree which

signifies the limit of resolution. Secondly, the logic gates establish a relationship among the basic

events that is required for the occurrence of a higher event. The higher event being the output of

the gate, and the lower events being the inputs of the gate. The gate symbol denotes the type of

relationship of the input events required for the output event.

3.1.1. Qualitative Evaluation of Fault Trees

The qualitative evaluations transform the FT logic into logically equivalent forms that provide

more focused information [Stamatelatos et al. (2002)]. There exist several methods for qualitative

analysis of standard fault trees with the same objective of determining the MCS that establish a

direct relationship between BEs and a predefined top event [Vesely et al. (1981)]. Qualitative

analysis can be performed using inductive (Bottom-Up) as well as deductive approach (Top-

Down) once the tree is translated to its equivalent Boolean equations. Before going ahead with

detail discussion of available algorithm for qualitative and quantitative analysis, it is necessary to

depict some logic rules and Boolean algebra utilized in the algorithms (See Table 4).

Table 4: Boolean algebra binary reduction rules

Mathematical symbolism Laws

∨ Logic OR gate

∧ Logic AND gate

𝐴 ∨ 𝐵 = 𝐵 ∨ 𝐴 Commutative Law

𝐴 ∨ 𝐴 = 𝐴 Idempotent Law

𝐴 ∧ (𝐵 ∨ 𝐶) = (𝐴 ∧ 𝐵) ∨ (𝐴 ∧ 𝐶) Distributive Law

𝐴 ∧ (𝐵 ∧ 𝐶) = (𝐴 ∧ 𝐵) ∧ 𝐶 Associative Law

𝐴 ∧ (𝐴 ∨ 𝐵) = 𝐴 Law of Absorption

(𝐴 ∨ 𝐵)′ = 𝐴′ ∨ 𝐵′ De Morgan’s theorem

The MCSs for a top event can be obtained using the well-known and the most common ‘Top-

Down’ approach. The approach can be implemented with the method of obtaining cut-sets

Page 43: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

27

(MOCUS) algorithm [Vesely et al. (1981); Kumamoto et al. (2000)]. It is based on the observation

that logic OR gates increase the number of cut-sets, whereas logic AND gates enlarge the size of

the cut-sets. MOCUS algorithm is demonstrated using the example FT.

Step-1: Naming logic gates

The logic gates are named as shown in the FT which includes, TOP, G1, G2, G3 and G4 gate.

Step 2: Numbering basic events

All basic events in the FT is named which includes; BE1, BE2, BE3, BE4, BE5 and BE6.

Step 3: Expansion of TOP

The uppermost logic gate “TOP” is identified, which forms the first element of the matrix. This

is an ‘OR’ gate, and thus “TOP” is replaced by a vertical array of the input (BE1 and G1) to the

gate:

TOP BE1

G1

Here, only G1 can be further expanded since BE1 is already a basic event.

Step 4: Expansion of G1

It may be noticed that G1 is a logic ‘AND’ gate, and thus G1 is replaced by a horizontal array of

input (G2 and G3) to the gate:

BE1

G2, G3

Step 5: Expansion of G2

Notice that G2 is and OR gate and thus it is replaced by a vertical array of the input (BE2 and

BE3) to the gate:

BE1

BE2, G3

BE3, G3

Here, BE2 and BE3 are basic events and cannot be further expanded. Only G3 can be expanded.

Step 5: Expansion of G3

Page 44: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

28

The gate G3 is a logic OR gate, and thus it is replaced by a vertical array of the input (BE4 and

G4) to the gate:

BE1

BE2, BE4

BE2, G4

BE3, BE4

BE3, G4

Step 5: Expansion of G4

The gate G4 is a logic OR gate, and thus it is replaced by a vertical array of the input (BE5 and

BE6) to the gate:

BE1

BE2, BE4

BE2, BE5

BE2, BE6

BE3, BE4

BE3, BE5

BE3, BE6

The above seven (7) rows in Step-5 represents seven (7) cut-sets for the FT. Notice that there are

no supersets (i.e., cut-sets that contains other complete cut-sets) in the cut-sets, and hence the 7

cut-sets are the minimal cut-sets of the tree.

{BE1}, {BE2, BE4}, {BE2, BE5}, {BE2, BE6}, {BE3, BE4}, {BE3, BE5}, and {BE3, BE6}

3.1.2. Quantitative Evaluation of Fault Trees

The quantification of FT top event (system unavailability/unreliability) can be performed given

the probability of occurrence of the individual BEs. MCS generated from qualitative analysis is

used for the quantification of top events. There are several methods for computation of the top

event such as the structure function method, rare event approximation method and the MCS upper

bound method, Beeson (2002). For illustration purpose assume the following failure probability

of BEs: BE1= 5 × 10−6, BE2= 3 × 10−5, BE3= 1 × 10−4, BE4= 1 × 10−4, BE5= 1 × 10−6,

and BE6= 5 × 10−6. Generally, approximation methods are used for the computation of top

event probability because the probabilities are relatively small, and computation of the exact

Page 45: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

29

probability becomes complex due to the dependencies between cut-sets. In this thesis, the MCS

upper bound method is used due to its simplicity and availability of software (CAFTA) that uses

the methodology for top event probability computation. A system failure occurs if at least one of

the MCS exist, and hence:

𝑃𝑟(Systemfailure) = 𝑃𝑟(AtleastonceMCSexists)

Thus;

𝑃𝑟 (𝑇𝑂𝑃) = 1 − (1 − 𝑃(𝑀𝐶𝑆𝑖))

𝑛

𝑖=1

(3-1)

Where, 𝑃(𝑀𝐶𝑆𝑖) = The probability of the 𝑖𝑡ℎ cut-set; and 𝑛 = Total number of cut-sets

Computation of the top event for the example FT using the above approximation method gives:

𝑃𝑟(𝑇𝑂𝑃) = 5.013 × 10−6

Furthermore, the relative importance of any MCSs can be obtained by taking the ratio of the MCS

probability to the total system probability. For example, taking the first MCS (a first order) of the

example system, the relative importance is determined to be 𝐼𝑚𝑚𝑐𝑠1 = 99.73%, which implies

the BE1 is the most critical component in the system.

Event Tree Analysis

Event tree analysis (ETA) is a well-established technique [WASH-1400 (1975); NUREG-1150]

that model an integrated overall plant response to an abnormal event. An event tree (ET) depicts

possible sequence of events logically and graphically following an initiating event (abnormal

condition that challenges the normal plant operation). The progression of an accident scenario

from an initiating event to some final plant state is systematically analyzed taking into account

the safety systems available and operator actions (success or failure) to counter or mitigate the

progression of an accident. ETA is an inductive technique that starts with an abnormal event and

progresses through the tree not only to determine the resulting consequences but also to enumerate

all possible accident scenarios. It should be noted here that ET technique is based on binary logic.

The accident sequence is characterized by the top events that generally represents the frontline

systems that is needed to respond to an initiating event. The process of obtaining these sequences

is called the qualitative analysis. Top events in an ET is typically modeled using FT technique.

Page 46: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

30

Within the current PRA framework, FTA and ETA are generally linked to determine the overall

plant core damage frequency, i.e., a FT model the individual systems and an ET model the overall

plant states [NUREG-1150; Kumamoto et al. (2000)]. Upon obtaining the top event probabilities

from FTs, accident sequences in the ET can be quantified.

For illustration purpose, consider a simple ET shown in Figure 3. The overall system consists of

three (3) safety systems (SS) that can either be functional or can be in a failed state. For instance,

these safety systems can be emergency core cooling system, electrical power, containment

system, etc. Given an initiating event (IE) with a probability 𝑃𝐼𝐸, the progression of the accident

scenario in the ET results in a total of 8 accident sequences. The sequences can be divided into

two general factors- the occurrence of IE and failure of frontline systems. Frontline systems in

this example are SS-1, 2 and 3, which can individually be modelled via FTs and linked to the ET.

Figure 2 FT which is considered as SS-1 is linked to the ET (Figure 3) for illustration purpose.

Figure 3: Example event tree

Assuming that the safety systems failure is independent from each other, accident sequences from

ET can simply be obtained by multiplying the probabilities of passing along each branch point on

any path through the tree by IE frequency. The accident sequences are shown in Table 5. For

instance, 𝑃𝐼𝐸 , 𝑃𝑆𝑆1, 𝑃𝑆𝑆2𝑎𝑛𝑑𝑃𝑆𝑆3 implies the probability of IE frequency, success probability of

safety system 1, 2 and 3 respectively, whereas ��𝑆𝑆1 implies SS-1 failure probability.

𝑃𝐼𝐸

SS-1 System State IE SS-2 SS-3

��𝑆𝑆1

𝑃𝑆𝑆1 𝑃𝑆𝑆2

𝑃𝑆𝑆2

��𝑆𝑆2

��𝑆𝑆2

𝑃𝑆𝑆3

𝑃𝑆𝑆3

𝑃𝑆𝑆3

𝑃𝑆𝑆3

��𝑆𝑆3

��𝑆𝑆3

��𝑆𝑆3

��𝑆𝑆3

Sequence-2

Sequence-1

Sequence-3

Sequence-5

Sequence-6

Sequence-7

Sequence-8

Sequence-4

Page 47: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

31

Table 5: Accident sequence from the event tree

Accident Sequence Probability

Sequence-1 𝑃𝐼𝐸𝑃𝑆𝑆1𝑃𝑆𝑆2𝑃𝑆𝑆3 0.998

Sequence-2 𝑃𝐼𝐸𝑃𝑆𝑆1𝑃𝑆𝑆2(1 − 𝑃𝑆𝑆3) 1.25E-07

Sequence-3 𝑃𝐼𝐸𝑃𝑆𝑆1(1 − 𝑃𝑆𝑆2)𝑃𝑆𝑆3 5.34E-07

Sequence-4 𝑃𝐼𝐸𝑃𝑆𝑆1(1 − 𝑃𝑆𝑆2)(1 − 𝑃𝑆𝑆3) 4.48E-11

Sequence-5 𝑃𝐼𝐸(1 − 𝑃𝑆𝑆1)𝑃𝑆𝑆2𝑃𝑆𝑆3 7.52E-09

Sequence-6 𝑃𝐼𝐸(1 − 𝑃𝑆𝑆1)𝑃𝑆𝑆2(1 − 𝑃𝑆𝑆3) 6.24E-13

Sequence-7 𝑃𝐼𝐸(1 − 𝑃𝑆𝑆1)(1 − 𝑃𝑆𝑆2)𝑃𝑆𝑆3 2.71E-12

Sequence-8 𝑃𝐼𝐸(1 − 𝑃𝑆𝑆1)(1 − 𝑃𝑆𝑆2)(1 − 𝑃𝑆𝑆3) 2.25E-16

If SS-1 failure probability (��𝑆𝑆1) is given by the top event of the FT in Figure 2, and assuming

that 𝑃𝐼𝐸 = 1.5 × 10−3, ��𝑆𝑆2 = 3.6 × 10−4, and ��𝑆𝑆3 = 8.3 × 10−5 respectively. ��𝑆𝑆2 and ��𝑆𝑆3

failure probability can be from another coupled FT similar to the 𝑃𝑆𝑆1 that was obtained from the

FT of SS-1. Frequency of each of the accident sequence can be computed with ease once the

failure probability of each of the frontline system (top event) is obtained. For example, the

accident sequence-1 frequency can be computed as:

Sequence-1 = 𝑃𝐼𝐸𝑃𝑆𝑆1𝑃𝑆𝑆2𝑃𝑆𝑆3 = 0.998

Similarly, the frequency of the rest of the sequences can be computed as shown in Table 5. Once

the individual sequences probability is calculated, the total system failure frequency can be

computed by summing all the accident sequences (AS) that lead to a system failure.

Total failure frequency = 𝐴𝑆𝑖

𝑛

𝑖=1

(3-2)

Where; 𝑛 = total number of accident sequences leading to a system failure.

The overall failure frequency of the system can be computed from the ET in Figure 3, and is

determined to be 6.72𝐸 − 07. Furthermore, assuming that a failure of SS-1 results in an overall

system failure, ET can be reduced from 8 to 5 accident sequences. The 5th accident sequence

frequency is: Sequence-5 = 𝑃𝐼𝐸��𝑆𝑆1 = 7.52𝐸 − 09.

Page 48: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

32

Dynamic Flowgraph Method

The DFM approach is based on representing the system of interest in a digraph (directed graph)

model, which is enriched with explicit identification of the cause-effect and time dependencies

among significant states of the system parameters that describe the system behavior [Garrett et

al. (1995); Guarro et al. (1996)]. DFM provides an integrated analytical framework for

systematically capturing the logical and dynamic behavior of a system respectively. The system

logical nature is expressed in terms of causal relationships between physical variables, whereas

dynamic behaviors are represented as a series of discrete state transitions. A system model can be

developed in DYMONDA platform with key system parameters that represents discrete-state

discrete-time system evolution. Once the DFM model is developed, the analysis can be performed

in an inductive or deductive manner. An inductive algorithm can be performed with a set of initial

and boundary conditions such as component states, time steps, direction rules, rate rules, etc., and

trace the model in the forward direction to obtain a sequence of events that follows from the initial

conditions. Whereas in the deductive algorithm, the user defines the desired specific system states

and time sequence, and the model is traced backward in time to identify the logical combination

of events leading to the defined top event.

The results obtained from an inductive or deductive algorithm are in the form of prime implicants

(PIs) that are analogous to the conventional FT MCS, with the exception that the PIs are time-

stamped. The probability of a PI occurrence can be calculated given the probability of each state

of a component/system parameter (user defined). Then, the quantification of the PIs can be

performed to obtain the exact top event probability, or the probability of system being in a specific

state sought by the analyst. Once a DFM model is built, it can be used repeatedly to produce MVL

and time-dependent PIs for a large number of possible top events that may be postulated for the

system. In this thesis, the methodology is implemented using the software DYMONDA

developed by ASCA Inc. A detail theoretical description, modelling process and analysis is

discussed in the subsequent sections.

3.3.1. Theoretical Basis

The DFM can be thought of as a series of snapshots of the classical FT technique, and hence the

methodologies share several similarities despite their differences. Similar to the FTA, the

Page 49: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

33

objective of DFM is to generate the PIs (MVL analogous of MCS encountered in binary FT) of a

desired top event. Some definitions are vital for the purpose of theoretical discussion.

Conjunction: The logical ‘OR’ operator (+,∪, or ∨)

Disjunction: The logical ‘AND’ operator (. ,∩, or ∧)

Literals: A primary event taking one of its state, e.g., event 𝐴1 (event A taking state

1 from 𝑛 number of possible states)

Monomial: A conjunction of literals, e.g., 𝑋 = 𝐴1𝐵2𝐶3 (a monomial 𝑋 is a

conjunction of event A, B and C in state 1, 2 and 3 respectively)

Top event: A disjunction of monomials

Implicant: A monomial 𝑋 of disjunctive form of a top event, such that 𝑇𝑂𝑃 ∩ 𝑋 =

𝑋 (MVL analogous to cut-set)

Prime Implicant: X is an implicant of a top event, and any other monomial Y subsumed by

X is not an implicant of the top event (MVL analogous to MCS)

Base of top event: Any disjunction of prime implicants which is equivalent to the top

function

Irredundant Base: A base which ceases to be a base if one of its prime implicants is removed

Complete Base: The disjunction of all prime implicants

Prime Implicants:

The prime implicants in DFM can be computed from a deductive as well as inductive analysis. In

this thesis, the deductive algorithm is emphasized due to its similarity with deductive FTA, which

makes it easier for comparison purpose. A deductive analysis starts with a defined top event and

track causality in reverse to uncover the root conditions that can cause the top event. As discussed

earlier, in the classical FTA, once a tree is developed, Boolean algebra can be applied to reduce

the tree to a unique disjunctive normal form in terms of its MCS. The MCSs are computed as a

conjunction of primary events [NUREG-6942]:

𝑀𝐶𝑆𝑗 = 𝑋𝑖(𝑗 )

𝑛

𝑖=1

(3-3)

Where, 𝑀𝐶𝑆𝑗 = Indicator variable for the 𝑗𝑡ℎ minimal cut-set

Page 50: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

34

𝑋𝑖(𝑗)

= Indicator variable for the 𝑖𝑡ℎ primary event in the 𝑗𝑡ℎ minimal cut-set

𝑛 = Number of primary events in the 𝑗𝑡ℎ minimal cut-set

The indicator variable for the top event (condition of interest- fail/success) can then be expressed

in disjunctive form as:

𝑋𝑇𝑂𝑃 = 1 − (1 −𝑀𝐶𝑆𝑗 )

𝑚

𝑗=1

(3-4)

where, 𝑚 = Total number of PIs.

The set of 𝑛 PIs obtained from a deductive analysis for a top event of interest, shown in Equation

(3-5), is first converted into a set of 𝑚 mutually exclusive implicants (MEIs) shown in Equation

(3-6). These MEIs can be thought of as the MVL equivalent of cut-sets that do not yield any cross

product term. The sum of the probabilities of these MEIs yields the exact probability of the top

event, as shown in Equation (3-7): Yau et al. (2007)

Top Event = 𝑃𝐼1⋁𝑃𝐼2⋁…𝑃𝐼𝑛 (3-5)

𝑃𝐼𝑖 ∉ 𝑃𝐼𝑗; 𝑓𝑜𝑟𝑎𝑛𝑦𝑖 ≠ 𝑗

Top Event = 𝑀𝐸𝐼1⋁𝑀𝐸𝐼2⋁…𝑀𝐸𝐼𝑚 (3-6)

Where; 𝑀𝐸𝐼𝑖 ⋀𝑀𝐸𝐼𝑗 = 0; 𝑓𝑜𝑟𝑖 ≠ 𝑗

𝑃(TopEvent) = 𝑃(𝑀𝐸𝐼1) + 𝑃(𝑀𝐸𝐼2) +⋯+ 𝑃(𝑀𝐸𝐼𝑚) (3-7)

This way the probability of the top event occurrence can be computed with the knowledge of the

probability of the primary events that constitutes the prime implicants. All PIs are time-stamped

and are referred to as “timed prime implicants”. For instance, 𝐴𝑖@𝑡 = 0 implies a component 𝐴

is in state 𝑖 at time 𝑡 = 0. PIs identified in a DFM analysis are conjunctions of primary events

that is sufficient to cause the top event but does not contain any shorter conjunction of events that

is sufficient to cause the top event. The base of a top event is first determined, which are of two

distinct types i.e., the irredundant base (IB) and complete base (CB) are more difficult to obtain

as compared to the FTA. Only the CB are unique and finite, Philip (2016).

Page 51: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

35

3.3.2. Method of Generalized Consensus

The determination of CB of a top event is of prime importance in the methodology. Several

methods exist for the determination of CB which include; Tabular method, Nelson method and

the Method of Generalized Consensus [Garribba et al. (1985); Philip (2016)]. In this thesis, the

CB is computed using the Method of Generalized Consensus due to the availability of well-

developed software code, which enables one to implement the methodology to a more complex

and realistic system. Quine developed the method of consensus for binary logic, which was then

extended by Yau (1997) to treat MVL. The methodology starts at the component level/basic event

having multiple states, and decision tables that establish a relationship among these components.

The decision table maps the input to the output parameter and can be constructed in an inductive

as well as deductive manner. A detail iterative process for the development of decision tables is

provided in the literature Salem et al. (1976; 1977; 1979). These decision tables are then merged

into a single table called the ‘critical transition table’. Several logic rules are applied on the critical

transition table to obtain the PIs [Ogunbiyi et al. (1981); Yau (1997)]. Thus, the determination of

PIs using the Generalized Consensus method consist of two iterative steps: Garribba et al. (1985)

a) Reduction: Application of multi-valued logic reduction rules including absorption,

reduction, merging and reduction-merging operation;

b) Development: Addition of monomials to the reduced set of implicants to obtain the

consensus terms (can be an implicant or PI).

The above two steps are repeated in an iterative manner until no more new consensus terms are

generated, and the total set of PIs include the reduced implicants from the original critical

transition table as well as the new PIs generated, [Garribba et al. (1985); Philip (2016)].

Decision Table Development

Decision tables are used to describe each possible output state of a component/sub-system as a

set of combinations of states of inputs, i.e., it maps the input parameters to the output. The use of

decision tables for FT construction have been studied comprehensively in [Salem et al. (1976;

1977; 1979)]. Decision tables can be constructed via inductive or deductive algorithm. The same

concept is implemented in DFM with the exception that a decision tables maps the state variables

as well as multistate system parameters. Detailed relationship between the multi-state nodes are

represented in decision tables. For example, if a sensor can be in three states (See Table 6) and is

Page 52: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

36

measuring a liquid level in a tank that can be at 0%, 50% and 100% (See Table 7), the measured

level which is the output of the sensor is clearly dependent on the sensor state and the liquid level.

The mapping of the input (liquid level- L and sensor state- SS) to output (measured level- ML)

can be performed as shown in Table 8. Of course, the measure level (ML) will have the same

number of states as the liquid level (L).

Table 6: Possible sensor states

No. Sensor State (SS) Representation

1. Normal 0

2. Failed Low -1

3. Failed High +1

Table 7: Liquid level in the tank and measured level

No. Liquid level State (L or ML) Representation

1. Level at 0 % 0

2. Level at 50 % 1

3. Level at 100 % 2

The decision table that maps the input to output is given in Table 8.

Table 8: Output decision table- the measured liquid level

No. Input Output

SS @t=-1 L @t=-1 ML @t=-1

1. 0 0 0

2. 0 1 1

3. 0 2 2

4. -1 - 0

5. +1 - 2

The above table can be interpreted as, if the level sensor is in normal condition, it will indicate

liquid level values as the original level in the tank. However, if the sensor is failed low or high, it

will indicate a low level (0 %) or high level (100 %) respectively no matter what the liquid level

is currently in, i.e., a “Don’t care condition”.

Page 53: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

37

A. Multi-Valued Logic Operations

The PIs of a top event can be obtained by merging the decision tables of individual constituents

in the system [Ogunbiyi et al. (1981)]. The merging operation is dependent on the structure of

the DFM model and the top event sought by the analyst, Philip (2016). The MVL operation

consist of absorption, merging, reduction and reduction-merging that are implemented to simply

the merged critical transition table, and eventually to determine the PIs of a system top event

sought by the analyst. The consensus operation is then applied to identify other PIs from the

irredundant base and is a vital step in obtaining the complete base of the top event [Yau (1997)].

1. Absorption Rule

A monomial X is said to subsume (absorb) a second monomial Y in a decision table if both the

monomials have identical output, and every input event in X also occurs in Y [Ogunbiyi et al.

(1981)]. This is equivalent to the absorption law in Boolean algebra. For example, if A, B and C

are three primary variables with multiple states, and:

𝑋 = 𝐴1. 𝐵2. 𝐶2 and 𝑌 = 𝐴1. 𝐶2 (3-8)

Then;

𝑋 + 𝑌 = 𝐴1. 𝐵2. 𝐶2 + 𝐴1. 𝐶2 = 𝐴1. 𝐶2 = 𝑌 (3-9)

Here, event A is in state 1, event B is in state 2 and event C is in state 2. The absorption rule

implemented in the transition table can be shown as:

Monomials Primary events Primary events

A B C

A C

X 1 2 2 1 2

Y 1 − 2

i.e., monomial X is absorbed into Y.

2. Merging

In a set of monomials 𝑋1, 𝑋2, … , 𝑋𝑛, if each 𝑋𝑖 contains identical literals except for one, and the

different literal contains a primary variable that is enumerated in all its states in 𝑋1, 𝑋2, … , 𝑋𝑛,

then the set of monomials can be merged into a single monomial [Yau (1997)]. For example, if

Page 54: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

38

A, B and C are primary variables, of which B has three distinct states 1, 2 and 3, and 𝑋1, 𝑋2 and

𝑋3 are monomials with:

𝑋1 𝑋2 𝑋3

𝐴1. 𝐵1. 𝐶2 𝐴1. 𝐵2. 𝐶2 𝐴1. 𝐵3. 𝐶2

It can be observed that event A and C remains the same, whereas event B is enumerated in all its

possible three states (1, 2, and 3). Thus monomials 𝑋1, 𝑋2 and 𝑋3 can be merged into a single

monomials. For example:

𝑋1 + 𝑋2 + 𝑋3 = 𝐴1. 𝐵1. 𝐶2 + 𝐴1. 𝐵2. 𝐶2 + 𝐴1. 𝐵3. 𝐶2 = 𝐴1. 𝐶2 (3-10)

The merging rule implemented in the transition table can be shown as:

Monomials Primary events Primary events

A B C

A C

𝑋1 1 1 2

1 2 𝑋2 1 2 2

𝑋3 1 3 2

i.e., three rows in the transition table have been merged into a single row.

3. Reduction

In a set of monomials 𝑋1, 𝑋2, … , 𝑋𝑛, if every literals of 𝑋𝑖 except one does not contradict with the

literals of the other 𝑋𝑗 (either 𝑋𝑗 contains the same literal, or 𝑋𝑗 does not have a literal with the

same primary event), and the contradicting literal contains a primary event that is enumerated in

all its states in 𝑋1, 𝑋2, … , 𝑋𝑛, then 𝑋𝑖 can be reduced by removing the contradicting literal

containing that primary event, Yau (1997). For example, if A, B and C are primary events, with

B having three distinct states 1, 2 and 3; and 𝑋1, 𝑋2 and 𝑋3 are monomials with:

𝑋1 𝑋2 𝑋3

𝐴1. 𝐵1. 𝐶2 𝐵2 𝐴1. 𝐵3

Page 55: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

39

The primary event A and C in 𝑋1 does not contradict with those in 𝑋2 and 𝑋3, and event B is

enumerated in all its three states in the 𝑋𝑖, then 𝑋1 can be reduced as:

𝑋1 + 𝑋2 + 𝑋3 = 𝐴1. 𝐵1. 𝐶2 + 𝐵2 + 𝐴1. 𝐵3 = 𝐴1. 𝐶2 + 𝐵2 + 𝐴1. 𝐵3 (3-11)

The reduction rule implemented in the transition table can be shown as:

Monomials Primary events

Primary events

A B C A B C

𝑋1 1 1 2 1 − 2

𝑋2 − 2 − − 2 −

𝑋3 1 3 − 1 3 −

It can be observed that 𝑋1 has been reduced from 𝐴1. 𝐵1. 𝐶2 to 𝐴1. 𝐶2.

4. Reduction-Merging

The reduction-merging rule is a combination of reduction and merging operation and is only

applicable to MVL systems. It is used to reduce an MVL decision table into its irredundant (most

reduced) form, Philip (2016). If a monomial obtained after reduction operation on a set of

monomials 𝑋1, 𝑋2, … , 𝑋𝑛, can absorb other monomials in the set, those subsuming monomials are

removed from the set. For example, if A, B and C are primary events, with event B having three

distinct states 1, 2 and 3, and 𝑋1, 𝑋2 and 𝑋3 are monomials with: [Yau (1997)]

𝑋1 𝑋2 𝑋3

𝐴1. 𝐵1. 𝐶2 𝐴1. 𝐵2. 𝐶2 𝐵3. 𝐶1

Then;

𝑋1 + 𝑋2 + 𝑋3 = 𝐴1 ∙ 𝐵1 ∙ 𝐶2 + 𝐴1 ∙ 𝐵2 ∙ 𝐶2 + 𝐵3 ∙ 𝐶1 (3-12)

= 𝐴1. 𝐶2 + 𝐴1. 𝐵2. 𝐶2 + 𝐵3. 𝐶1 (3-13)

𝑋1 + 𝑋2 + 𝑋3 = 𝐴1. 𝐶2 + 𝐵3. 𝐶1 (3-14)

i.e., 𝑋1 is first reduced from 𝐴1. 𝐵1. 𝐶2 to 𝐴1. 𝐶2, and then 𝑋2 (𝐴1. 𝐵2. 𝐶2) is absorbed into the new

monomial 𝐴1. 𝐶2. The reduction-merging rule implemented in the transition table is shown as:

Page 56: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

40

Monomials Primary events

Primary events

A B C A B C

𝑋1 1 1 2 1 − 2

𝑋2 1 2 2 − − −

𝑋3 − 3 1 − 3 1

5. The Consensus Operation

A critical transition table is not guaranteed to contain a complete set of PIs, particularly if

multistate components and success states exist. Hence to obtain a CB of a top event, the consensus

operation is performed that creates new terms out of the existing terms in the critical decision

table by mixing and matching their input events [Ogunbiyi et al. (1981)]. The consensus operation

can be illustrated as:

Given a primary event A having 𝑛 distinct state 1, 2, … , 𝑛, and monomials 𝑋1, 𝑋2, … , 𝑋𝑛 do not

contain the primary event A, then a consensus operation on the monomials 𝐴1 ∙ 𝑋1, 𝐴2 ∙

𝑋2, … , 𝐴𝑛 ∙ 𝑋𝑛 yields a new monomial 𝑋1 ∙ 𝑋2 ∙ … ∙ 𝑋𝑛 provided this new monomial is not

inherently false. This can be expressed as: [Yau (1997)]

𝐴1 ∙ 𝑋1 + 𝐴2 ∙ 𝑋2 +⋯+ 𝐴𝑛 ∙ 𝑋𝑛 = 𝐴1 ∙ 𝑋1 + 𝐴2 ∙ 𝑋2 +⋯+ 𝐴𝑛 ∙ 𝑋𝑛 + 𝑋1 ∙ 𝑋2 ∙ … ∙ 𝑋𝑛 (3-15)

It can be observed that the consensus operation yields a new implicant (𝑋1 ∙ 𝑋2 ∙ … ∙ 𝑋𝑛) from the

base 𝐴1 ∙ 𝑋1 + 𝐴2 ∙ 𝑋2 +⋯+ 𝐴𝑛 ∙ 𝑋𝑛.

For example, if A, B and C are primary events, with event A having two distinct states 1 and 2,

and 𝑋1 and 𝑋2 are monomials such that:

𝑋1 = 𝐵1 and 𝑋2 = 𝐶2 (3-16)

The consensus operation on the monomials 𝐴1. 𝐵1 and 𝐴2. 𝐶2 is given as:

𝐴1 ∙ 𝑋1 + 𝐴2 ∙ 𝑋2 = 𝐴1 ∙ 𝐵1 + 𝐴2 ∙ 𝐶2 (3-17)

= 𝐴1 ∙ 𝐵1 + 𝐴2 ∙ 𝐶2 + 𝐵1𝐶2 (3-18)

A new monomial 𝑋3 = 𝐵1𝐶2 is generated from the existing monomial 𝑋1 and 𝑋2. The consensus

operation implemented in the transition table is shown as:

Page 57: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

41

Monomials Primary events

Primary events

A B C A B C

𝑋1 1 1 − 1 1 −

𝑋2 2 − 2 2 − 2

𝑋3 − − − − 1 2

The consensus operation on the two monomials (two rows in the table) yields a third row in the

critical transition table or a third monomial 𝑋3.

3.3.3. DFM Modelling Process

A DFM model expresses the logical and dynamic behavior of a system with a network of nodes

created by discretizing the key system parameters and state variables. These system nodes are

then linked together to represent the cause-and-effect and time-dependent relationships which

exist among key system parameters. Decision tables within the nodes constructed from the

equations governing the system behavior represents the functional mappings among the system

parameters [Yau et al. (1998)]. The application of DFM within a PRA framework typically

involves three essential steps: [NUREG-6465]

1. Development of the DFM model for which safety analysis is desired

2. Analysis of DFM model (deductive or inductive)

3. Quantification of the prime implicants.

The DFM model encompasses all the key system parameters and are represented by a network of

nodes. The fundamental DFM modelling elements are shown in Figure 4. A DFM model utilize

these elements to represent the temporal and logical relations that exist among key system

parameters, Yau (1997).

Process variable nodes:

They represent physical variables that captures the essential functional behaviour, continuous or

discrete of a system. A variable represented by a process variable node is discretized into finite

number of intervals. The number of discretized intervals of a state variable is chosen on the basis

of balance between the accuracy of the model, and the complexity introduced by higher numbers

of variable states. For example, steam generator liquid level discretization.

Page 58: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

42

Process variable

Conditioning node

Transfer box

Transition box

Causality Edge

Conditioning Edge

Figure 4: Basic building blocks of DFM Model

Causality edge:

They are used to connect process variable nodes to indicate the existence of a cause-and-effect

relationship between the variables described by the nodes. Each causality edge is connected to

a transfer box to describe the precise functional relationships among the connected nodes.

Transfer box:

A transfer box represents a transfer function among system parameter nodes. A decision table is

associated with each transfer box and is used to quantify the relationships between its input and

output system parameters. A decision table is a multi-dimensional matrix whose dimension is

equal to one plus the number of its inputs.

Transition box:

A transition box is similar to transfer box with the exception that, a time lag or time transition is

assumed to occur between the time when the input variable become true and the time when the

output variable associated with the inputs is reached. This time delay is a characteristic of the

transition which is being modelled and is treated as an attribute of the transition box.

Conditioning Nodes:

It explicitly identifies component failure states, changes of process operation regimes and

modes, and switching actions. These nodes can affect the logic superstructure of a system by

modifying the causal relations between process variable nodes.

Conditioning Edge:

Page 59: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

43

They are used to represent true discrete behaviour in the system. Conditioning edge link

parameter nodes to transfer boxes, indicating the possibility of using a different transfer

function to map input variable into output variable states.

3.3.4. Timed-Fault Tree Construction

Timed-fault tree maybe viewed as sequences of static fault trees at different time steps

representing the evolution of logical combinations of events leading to a top event, Garrett et al.

(1995). The generation of a timed-fault tree is of prime importance due to the fact that most

existing NPP PRAs model are based on classical techniques. In order that the PIs generated from

the DFM model are incorporated into an existing plant PRA model, its conversion into FT become

necessary. Timed-fault trees can be constructed from a DFM model using by backtracking

through the decision tables. Throughout the backtracking process, the following consistency rules

are applied: [Yau (1997); NUREG/CR-6465]

a. Physical consistency rule

Physical consistency rules are applied to eliminate physically impossible conditions from the

timed fault trees. For instance, a state variable cannot be in two different states at the same time

step in the time FTs; example, a component cannot be in ‘ON’ and ‘OFF’ states in the same time

step. The physical consistency rule is similar to the consistency rules applied in the classical FT.

If a system parameter appears twice, in different states, in the same time step and under the same

logic AND gate, then all the parameters beneath the first logic AND gate above the second

occurrence of the event is pruned from the tree due to physical inconsistency. [Garrett et al.

(1995); Yau (1997)].

b. Dynamic consistency rule

Dynamic consistency rules are applied to timed FTs to eliminate branches which cannot occur

due to constraints from the dynamic behavior of the system under consideration. Dynamic rules

are developed from the analyst's knowledge and assumptions with regard to system dynamic

behavior. For instance, if the time step considered in a DFM model is 1 second, and water level

in a steam generator was initially at 0%, the water level in the next time step cannot increase to

80% (impossible condition due to system dynamics). Similar to physically consistency rule, the

dynamically inconsistent branches, including all of the sub-branches connected to it via the first

Page 60: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

44

parent AND gate are pruned. Further pruning is performed if eliminated branches cause other

events to become impossible [Garrett et al. (1995); Yau (1997)].

Classical Markov Chain

Markov chain enables one to model a sequence of random variables which correspond to the

states of a system, and the system state at any point in time is dependent only on the system state

in the previous time. Markov chain captures system dynamic nature with its two-fundamental

property of characterizing a process evolution by a set of distinct system states and transition

between these states. The mathematical model of a Markovian system is thus a collection of

system states and the conditional probabilities between these states. The system dynamics is

represented by a set of coupled linear differential equations (Chapman-Kolmogorov equation),

which can be solved using standard analytical methods.

3.4.1. Formulation of the Markov Chain

Consider a discrete-time (𝑛) stochastic process with a sequence of random variables

𝑋0, 𝑋1, 𝑋2, 𝑋3, … , 𝑋𝑛 (𝑋𝑛, 𝑛 ≥ 0) such that each random variable 𝑋𝑛 takes values in a discrete set

𝑆 (𝑆 = ℕ) i.e., the state space, then:

ℙ(𝑋𝑛+1 = 𝑗|𝑋𝑛 = 𝑖,𝑋𝑛−1 = 𝑖𝑛−1,… ,𝑋0 = 𝑖0) = ℙ(𝑋𝑛+1 = 𝑗|𝑋𝑛 = 𝑖)

(3-19)

∀𝑛 ≥ 0; 𝑗, 𝑖, 𝑖𝑛−1, …… , 𝑖0 ∈ 𝑆

It may be observed from the equation that the process is memoryless, and the next system state is

dependent only on the current state. Furthermore, 𝑋𝑛 is time homogeneous if:

ℙ(𝑋𝑛+1 = 𝑗|𝑋𝑛 = 𝑖) = (𝑋1 = 𝑗|𝑋0 = 𝑖) = 𝑃𝑖𝑗

(3-20)

i.e., the transition probabilities are independent of 𝑛. Thus, for an (𝑛 + 𝑚) time steps with 𝑛 ≥

0,𝑚 ≥ 0, the above equation become:

ℙ(𝑋𝑛+𝑚 = 𝑗,𝑋𝑛 = 𝑘|𝑋0 = 𝑖) = 𝑃𝑖𝑘𝑛𝑃𝑘𝑗

𝑚

(3-21)

Summing up all the 𝑘 yields the Chapman-Kolmogorov equation:

Page 61: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

45

𝑃𝑖𝑗𝑛+𝑚 = 𝑃𝑖𝑘

𝑛𝑃𝑘𝑗𝑚

𝑘∈ 𝑆

(3-22)

Also by necessary probability condition:

ℙ(𝑋1 = 𝑗|𝑋0 = 𝑖)

𝑗∈𝑆

= 1

Here, 𝑃𝑖𝑗 denotes the probability that the chain, whenever in state 𝑖 transits into the next state 𝑗;

and takes a value 0 ≤ 𝑃𝑖𝑗 ≤ 1 for all 𝑖 and 𝑗. For each 𝑖 ∈ 𝑆, the transition probability matrix

satisfies the following condition (rows of the transition matrix):

𝑃𝑖𝑗𝑗 ∈ 𝑆

= 1

(3-23)

3.4.2. Computation of State Probabilities

For a Markovian system with a finite number of discrete-state, the objective is to determine the

system state transition probabilities at a given point in time. The stochastic rate of transition

among system states is described by a set of ordinary linear differential equations. The state

probabilities can be determined by solving the set of coupled Chapman-Kolmogorov differential

equations. The general form of the first order linear differential equations is as follow:

𝑑𝑃𝑖(𝑡)

𝑑𝑡= − 𝑎𝑖𝑗𝑃𝑖(𝑡)

𝑛

𝑗=1𝑗≠𝑖

+ 𝑎𝑗𝑖𝑃𝑗 (𝑡)

𝑛

𝑗=1𝑗≠𝑖

(3-24)

Where; 𝑛 = total number of system states

𝑃𝑖(𝑡) = state probability for the 𝑖𝑡ℎ state

𝑎𝑖𝑗 = the rate of transition from state 𝑖 to state 𝑗 (typically component repair and failure rates)

The computation of state probabilities or solution of a Markov model involves three steps:

a) Model development: This step involves the development of state transition diagram such

as determination of system states, the possible transition between these states, and the

transition rates. It may also include labeling of system states in a qualitative manner such

as system operational, failure or partial failure.

Page 62: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

46

b) Generation of ODEs: The Markovian state transition diagram developed in the preceding

step is utilized to generate a set of linear ordinary differential equations that describes the

system behavior and characteristics.

c) Determination of State Probabilities: The obtained linear ODEs are solved to determine

the system states probabilities. Many standard techniques are available for solving the set

of ODEs, such as analytical methods, Laplace transform and numerical integration.

For illustration purpose on computation of state probabilities, two cases are considered: (1) Non-

repairable system; and (2) Repairable system.

1. A non-repairable system

Consider a non-repairable component with two states, i.e., normal and failed states. The system

can only make a transition from operational state to the failed state with a constant failure rate of

𝜆 due to its nature of non-reparability. The system can possibly take two states, and can be

represented in the transition diagram (Figure 5):

State 1: System operational (𝑆𝑜)

State 2: System failed, or absorbing state (𝑆1)

Figure 5: Markov state-transition diagram of a two state component

The system of linear ordinary differential equations (ODEs) from the state-transition diagram are:

𝑑𝑆0(𝑡)

𝑑𝑡= −𝜆𝑆0(𝑡)

(3-25)

𝑑𝑆1(𝑡)

𝑑𝑡= 𝜆𝑆0(𝑡)

(3-26)

With the assumption that the system is fully operational at 𝑡 = 0, 𝑆0(0) = 1 and 𝑆1(0) = 0,

solution of the differential equation, and probability density function of system failure is:

𝑆0(𝑡) = 𝑒−𝜆𝑡 and 𝑆1(𝑡) = (1 − 𝑒−𝜆𝑡)

Page 63: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

47

𝑓(𝑡) =𝑑𝑆1(𝑡)

𝑑𝑡=

𝑑

𝑑𝑡(1 − 𝑒−𝜆𝑡 ) = 𝜆𝑒−𝜆𝑡

(3-27)

From definition, the system reliability and unreliability are simply represented by 𝑆0(𝑡) and 𝑆1(𝑡).

For a component with a constant failure rate of 𝜆 = 0.005 failures per hr and a mission time of

1200 hours, the reliability and unreliability of the system is plotted in Figure 6.

Figure 6: System reliability and unreliability for a two state non-repairable system

2. A repairable system

Similar to the previous case, consider a system with two states i.e., operational and failed, with

the difference that the system is repairable. The system can make a transition from operational to

failed state with a rate 𝜆, and from failed state to operational state with a repair rate of 𝜇. Hence,

the probability that the system will remain in a given state is (1 − 𝜆) and (1 − 𝜇) for state 𝑆0(𝑡)

and 𝑆1(𝑡) respectively. The Markov state transition diagram is shown in Figure 7.

Figure 7: State-transition diagram of a two state component with repair

The linear differential equations describing the time-dependent system characteristics are:

Page 64: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

48

𝑑𝑆0(𝑡)

𝑑𝑡= (1 − 𝜆)𝑆0(𝑡) + 𝜇𝑆1(𝑡)

(3-28)

𝑑𝑆1(𝑡)

𝑑𝑡= 𝜆𝑆0(𝑡) + (1 − 𝜇)𝑆1(𝑡)

(3-29)

The solution of the above ODEs is:

𝑆0(𝑡) =𝜇

(𝜆+𝜇)+

𝜆

(𝜆+𝜇)𝑒−(𝜆+𝜇)𝑡 and 𝑆1(𝑡) =

𝜆

(𝜆+𝜇)(1 − 𝑒−(𝜆+𝜇)𝑡) (3-30)

For a component with a constant failure rate of 𝜆 = 0.005 failures per hr, repair rate of 𝜇 =

0.00125 repairs per hr and a mission time of 1500 hours, the reliability (𝑆0(𝑡)) and unreliability

(𝑆1(𝑡)) of the system is plotted in Figure 8.

Figure 8: State probabilities for component with repair

Cell-to-Cell Mapping Technique

Cell-to-Cell Mapping Technique (CCMT) is a systematic procedure to describe the dynamics of

both linear and nonlinear systems in discrete time and discretized system state-space, or the

subspace of the state variables only, Hsu (1980). The fundamental concept of the methodology is

to discretize the state space into a finite number of cells and determine the probability that the

system will occupy a discretized cell as time evolve by means of standard numerical integration

methods such as Runge-Kutta method. Thus, a state variable represents the dynamics of a system

Page 65: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

49

in discrete quantity that takes every possible value 𝑥 ∈ ℝ. Cell mapping technique can be

categorized as: simple cell mapping, generalized cell mapping and interpolated cell mapping,

Spek (1994). In this thesis, generalized cell mapping is used for benchmark system analysis, due

to its efficiency and accuracy of the methodology to predict the long term dynamic behavior of

linear and non-linear systems [Hsu (1980); Spek (1994)]. CCMT defines the system states in

terms of both system configurations i.e., vectors of the discrete states occupied by the system

components, and cells occupied by state variables. This allows modelling system configuration

(instantaneous) changes upon crossings of threshold values (e.g., a valve closing when the liquid

level exceeds a pre-set value). Consider a simple dynamic system governed by the ODE:

𝑑𝑥

𝑑𝑡= 𝑓(𝑥,𝛼, 𝑡)

(3-31)

Where, 𝑥 is an 𝑁 dimensional state variable vector in ℝ𝑁, 𝛼 is the 𝐾 dimensional system

parameter vector, 𝑡 is the time variable, and 𝑓 is a non-linear function vector. Equation (3-22) can

be integrated over one period to relate the state of the system at the end of one period to the state

at the end of the next period.

𝑥(𝑡 + ∆𝑡) = 𝑥(𝑡) + 𝑓(𝑥,𝛼)𝑑𝑡

𝑡+∆𝑡

𝑡

(3-32)

The above solution for 𝑥 in discrete point mapping form can be written as:

𝑥(𝑛 + 1) = 𝐺(𝑥(𝑛), 𝛼) (3-33)

Here, a point 𝑥(𝑛) in the state space is mapped by the mapping transition probability matrix (𝐺)

after one discrete time step into a point 𝑥(𝑛 + 1), and hence the name point mapping method.

This enables one to represent the system dynamics in finite sequence of discrete system states.

This very feature of point mapping of the system dynamics as a continuum of dimension N state

space results in poor computational efficiency. This leads to the idea of considering a state

variable not as a continuum of points but as a collection of finite number of cells, Hsu (1980).

The region of interest Ω in the state space is chosen and partitioned into finite number of subsets

𝐽𝑖 ∈ Ω. Each subset 𝐽𝑖 is considered as the smallest entity in the control state space and called a

regular or computational cell. All the cells can be labeled with an integer 𝑧, from 1 to 𝑁𝑐. The

Page 66: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

50

region outside the control space 𝛤 can also be divided into finite number of cells called the sink

cells, from 1 to ��𝑠.

The computational cells can be discretized or partitioned in several ways and is dependent on the

analyst’s choice. For example, the coordinate axis of a state variable 𝑥𝑖 (𝑖 = 1, 2, … ,𝑁) can be

divided into a finite number of intervals with an interval size of ℎ𝑖. The interval 𝑍𝑖 of the 𝑥𝑖 axis

is defined to be one which contains all 𝑥𝑖 satisfying: [Hsu (1980)]

(𝑍𝑖 − 1/2)ℎ𝑖 ≤ 𝑥𝑖 < (𝑍𝑖 + 1/2)ℎ𝑖

ℎ𝑖 =𝑥𝑖(𝑚𝑎𝑥)

− 𝑥𝑖(𝑚𝑖𝑛)

𝑀𝑖, 𝑖 = 1, 2,… ,𝑁

(3-34)

Where, 𝑍𝑖 is an integer, and 𝑀𝑖 is the number of intervals. A 𝑁-tuple 𝑍𝑖 , 𝑖 = 1, 2, … ,𝑁 is called a

cell vector of the state space and is denoted by ℤ. The collection of the regular cells and the sink

cell forms a cell state space 𝑆 = {0, 1, … , 𝑁𝑐}. This way the continuous state space is transformed

into a 1-dimensional integer array. Figure 9 shows an example state space discretization scheme

with 𝑁𝑐 computational cells in the control space, and ��𝑠 sink cell in 𝛤.

Figure 9: A two dimensional state space discretization

𝐽 ҧ3

0

𝐽2 𝐽1 𝐽3

State Variable (𝑥2)

𝛼1(𝑢)

𝛼2(𝑙)

𝛼1(𝑙)

𝛼2(𝑢)

State Variable (𝑥1)

Ω

𝑁𝑐 𝑁𝑐

𝐽 ҧ1 𝐽 ҧ2

𝑁𝑠 Sink cells

Regular cells

Page 67: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

51

A point 𝑥(𝑥𝑖, 𝑖 = 1, 2, … , 𝑁) belongs to a cell ℤ(𝑍𝑖 = 1, 2, … ,𝑁) if 𝑥𝑖 belongs to 𝑍𝑖 for all 𝑖. By

an appropriate application of rules to identify 𝑥(𝑛) and 𝑥(𝑛 + 1) with corresponding

computational cells 𝑍(𝑛) and 𝑍(𝑛 + 1), one can associate to the point-to-point mapping

(Equation 3-24) a cell-to-cell mapping ℂ. Thus, cell mapping is derived from point mapping by

appropriate discretization process.

ℤ(𝑛 + 1) = ℂ(ℤ(𝑛),𝛼)

ℤ𝑖(𝑛 + 1) = ℂ𝑖(ℤ(𝑛),𝛼)

(3-35)

Where ℂ maps a set of integers to a set of integers, and thus ℂ is an integer-valued cell mapping.

Note that, it is not necessary for all the computational cells to be identical, rather a computational

cell can be of any shape. However, rectangle is the most commonly used shape because of its

convenience, Chen (2004). Further Hsu (1981) developed the generalized cell mapping theory by

removing the restriction in simple cells mapping that a computational cell ℤ(𝑛) is mapped by ℂ

into only a single cell ℤ(𝑛 + 1). The generalized theory enables a computational cell ℤ(𝑛) to be

mapped into several image cells, with each of the image cells having a definite fraction of the

total probability. For instance, if a system is in a computational cell ℤ(𝑛) at time 𝑡 = 𝑛, the

evolution of the system state in the next time step (𝑛 + 1) is given by:

ℤ(1)(𝑛 + 1) = 𝑝(1), ℤ(2)(𝑛 + 1) = 𝑝(2), … , ℤ(𝑖)(𝑛 + 1) = 𝑝(𝑖); (𝑖 = 1, 2, … ,𝑁)

Here, 𝑝(1) implies the probability of mapping ℤ(𝑛) into cell (1), and 𝑝(2) implies the probability

of mapping ℤ(𝑛) into cell (2), and so forth. i.e., a single computational cell can be mapped into

several other cells, or even into itself with some cell-to-cell transition probabilities. Of course, by

definition and probability theory:

𝑝(𝑖)

𝑖

= 1

(3-36)

i.e., the sum parameter covers all the possible image cell that a computational cell can take in the

next time step. This allows one to describe the system state at any point in time with adequate

probabilities of finding a system in several cells. Let:

𝑆 = Closed set of cells of interest (cells within the control space)

Page 68: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

52

𝜓𝑖(𝑛) = the probability of the state of the system being in cell 𝑖 at 𝑡 = 𝑛

The vector 𝝍(𝑛) with component 𝜓𝑖(𝑛), 𝑖 = 1, 2, … , 𝑁 is called the cell probability vector.

𝑝𝑖𝑗 = Probability of cell 𝑗 being mapped into cell 𝑖 in one mapping time step

𝑃 = Transition probability matrix with components 𝑝𝑖𝑗

The following properties of 𝜓𝑖(𝑛) and 𝑝𝑖𝑗 is inferred from the above, as is given by:

𝜓𝑖(𝑛)

𝑖∈𝑆

= 1

(3-37)

𝑝𝑖𝑗𝑖∈𝑆

= 1

(3-38)

𝜓𝑖(𝑛) ≥ 0 and 𝑝𝑖𝑗 ≥ 0

Thus, the system evolution can be represented by:

𝝍(𝑛 + 1) = 𝑃𝝍(𝑛)

(3-39)

For a given initial cell probability vector 𝝍(0) at 𝑛 = 0, the subsequent system evolution is

simply given by:

𝝍(𝑛) = 𝑃𝑛𝝍(0)

(3-40)

The mapping probability matrix, 𝑃 completely describe the system evolution and dynamics in the

cell space. Thus, useful information of the system can be obtained by solving the transition

mapping probability matrix 𝑃. It can be observed that the above equations transformed the system

dynamics into a finite state Markov model with stochastic matrix 𝑃 governing the system

dynamics.

Coupled Markov-CCMT Model

The coupled Markov-CCMT model integrates the classical Markov model with CCMT to

represent the possible coupling between failure events that can originate from the dynamic

interactions between system components and controlled state variables, and among the different

Page 69: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

53

constituents of the system [NUREG-6985]. Figure 10 provide the framework and methodology

implementation flowchart.

Figure 10: Coupled Markov-CCMT flowchart for integrated system analysis

The coupled Markov-CCMT models a system evolution in discrete time through the probability

𝑃𝑛,𝑗(𝑡) that the state variables are in a predefined regions or cells 𝑉𝑗 in the state space at time

(𝑘 + 1)∆𝑡 with the components state combination 𝑛 (e.g., valve failed-closed), given that the

state variables was in cell 𝑉𝑗′ at time 𝑘∆𝑡 with a component state combination 𝑛′, Hassan and

Aldemir (1990). The dynamic behavior of the system is usually described by a set of differential

or algebraic equations, as well as the set of control laws.

1. Model Assumptions

The methodology is based on the following assumptions [Aldemir (1987); Hassan and Aldemir

(1990); Belhadj and Aldemir (1992)]:

1. The system configuration or components state do not change during the time interval

[𝑘∆𝑡, (𝑘 + 1)∆𝑡] but possibly at (𝑘 + 1)∆𝑡;

System Description

System Analysis

Analysis of Type-I

Interactions

Analysis of Type-II

Interactions

Dynamic and Control

Laws

Cell-to-cell Mapping

Technique

Finite States

Description

Markov Model

System Modelling

(Deterministic)

Stochastic System

Model (Probabilistic)

Integrated Deterministic and

Probabilistic Analysis

Page 70: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

54

2. For a given component state combination 𝑛 and cell 𝑉𝑗, 𝑃𝑛,𝑗(𝑘∆𝑡) is uniformly distributed

over 𝑉𝑗;

3. If the modeling is conducted in the state variable state space, no two controlled variable

trajectories arrive at the same point in state space at the same time and move in different

directions for the same component state combinations.

Assumptions 1 and 2 lead to an approximation of the probabilistic system dynamics. Assumption

1 also leads to an approximation of the failure characteristics of the components. Under these

assumptions, the probabilistic evolution of the system in time 𝑃𝑛,𝑗(𝑘 + 1)∆𝑡 can be determined

recursively following an inductive Markov-CCMT implementation: [Aldemir (1987); Hassan

and Aldemir (1990)]

𝑃𝑛 ,𝑗 ((𝑘 + 1)∆𝑡) = 𝑔(𝑗 𝑗′⁄ ,𝑛′ ,𝑘∆𝑡).ℎ(𝑛 𝑛′⁄ , 𝑗′ → 𝑗,𝑘∆𝑡).𝑃𝑛 ′ ,𝑗 ′ (𝑘∆𝑡)

𝐽

𝑗 ′ =1

𝑁

𝑛 ′ =1

(3-41)

Where;

𝑞𝑛 ,𝑗𝑛 ′ ,𝑗 ′ (𝑘∆𝑡) = 𝑔(𝑗 𝑗′⁄ ,𝑛′ , 𝑘∆𝑡).ℎ(𝑛 𝑛′⁄ , 𝑗′ → 𝑗, 𝑘∆𝑡)

(3-42)

Therefore;

𝑃𝑛 ,𝑗 ((𝑘 + 1)∆𝑡) = 𝑞𝑛 ,𝑗𝑛 ′ ,𝑗 ′ (𝑘∆𝑡).𝑃𝑛 ′ ,𝑗 ′ (𝑘∆𝑡)

𝐽

𝑗 ′ =1

𝑁

𝑛 ′ =1

(3-43)

The above equation is equivalent to the Chapman–Kolmogorov equation in discretized state space

and discretized time. Here;

∆𝑡 = Cell-to-cell mapping time step

𝑃𝑛′,𝑗′(𝑘∆𝑡) = Pr{Controlled variables are in cell 𝑗′, and component state combination is in state

𝑛′ at time 𝑡 = 𝑘∆𝑡

𝑃𝑛,𝑗((𝑘 + 1)∆𝑡) = Pr{Controlled variables are in cell 𝑗, and component state combination is in

state 𝑛 at time 𝑡 = (𝑘 + 1)∆𝑡

𝑔(𝑗 𝑗′⁄ , 𝑛′, 𝑘∆𝑡) = Pr{Controlled variables are in cell 𝑗 at time (𝑘 + 1)∆𝑡 given that the

controlled variables are in cell 𝑗′ at time 𝑡

Page 71: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

55

ℎ(𝑛 𝑛′⁄ , 𝑗′ → 𝑗, 𝑘∆𝑡) = Pr{Component state combination is in state 𝑛 at time (𝑘 + 1)∆𝑡 given

that the component state combination is in state 𝑛′ at time 𝑡, and the state variables move from

cell 𝑗′ to cell 𝑗 during [𝑘∆𝑡, (𝑘 + 1)∆𝑡].

𝑞𝑛,𝑗𝑛′,𝑗′

= Elements of the transition matrix for the Markov model

Since, 𝑉𝑗 cover the whole control variable state space and 𝑁 includes all possible component state

combinations, by essence of probability theory:

𝑃𝑛 ,𝑗 (𝑡)

𝐽

𝑗=1

𝑁

𝑛=1

= 1

(3-44)

𝑞𝑛 ,𝑗𝑛 ′ ,𝑗 ′

(𝑘∆𝑡)

𝐽

𝑗 ′=1

𝑁

𝑛′=1

= 1

(3-45)

The parameter 𝑞𝑛,𝑗𝑛′,𝑗′

is a conditional probability that accounts for the simultaneous occurrence,

and interactions between the state variables and the system components which are statistically

dependent. Hence, to compute 𝑞𝑛,𝑗𝑛′,𝑗′

, the conditional probability between 𝑔 and ℎ have to be

known. However, neither 𝑔 or ℎ can be determined individually due to the statistical dependency

Aldemir (1987). To overcome this drawback, the methodology approximates 𝑔 using Assumption

1 that the system state does not change during the short interval [𝑡, (𝑡 + 𝑘∆𝑡)] but can

simultaneously change during (𝑡 + 𝑘∆𝑡).

2. Determination of cell-to-cell transition probabilities, 𝑔(𝑗 𝑗′⁄ , 𝑛′, 𝑘∆𝑡)

The cell-to-cell transition probabilities depends on the system dynamics, the deterministic control

laws, and the possible system configurations. 𝑔(𝑗 𝑗′⁄ , 𝑛′, 𝑘∆𝑡) represents the system dynamics

under normal and failed state of the components/sub-systems, i.e., conditional probability that the

state variables are in cell 𝑉𝑗 at time 𝑡 = (𝑘 + 1)∆𝑡 given that:

• The state variables are in cell 𝑉𝑗′ at time 𝑡 = 𝑘∆𝑡

• The component state combination is in 𝑛′ at time 𝑡.

The 𝑔(𝑗 𝑗′⁄ , 𝑛′, 𝑘∆𝑡) can be found from [Hassan and Aldemir (1990); Belhadj et al. (1992)]:

Page 72: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

56

𝑔(𝑗 𝑗′⁄ , 𝑛′ ,𝑘∆𝑡) =

𝑑𝑥′

𝑣𝑗 ′𝑒𝑗 ��(𝑘+1)∆𝑡(𝑥

′ ,𝛼𝑛 ′ ,𝑘∆𝑡) ; 𝑖𝑓 𝑉𝑗 ′ ∈ 𝑅𝑉𝑗 ′

𝛿𝑗 ,𝑗 ′ 𝑂𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒

(3-46)

Where;

𝛿𝑗 ,𝑗 ′ = 𝐾𝑟𝑜𝑛𝑒𝑐𝑘𝑒𝑟 𝑑𝑒𝑙𝑡𝑎 = 1 ; 𝑖𝑓 𝑗′ = 𝑗0 𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒

(3-47)

𝑒𝑗 (𝑥) = 𝑆𝑡𝑒𝑝 𝑓𝑢𝑛𝑐𝑡𝑖𝑜𝑛 = 1 ; 𝑖𝑓 ��(𝑘+1)∆𝑡 ∈ 𝑉𝑗0 𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒

(3-48)

Where; 𝑥′ = Initial value of the state variable at time 𝑡 = 𝑘∆𝑡

��(𝑘+1)∆𝑡(𝑥′, 𝛼𝑛′ , 𝑘∆𝑡) = The location of the system in the state variable space at time 𝑡 = (𝑘 +

1)∆𝑡, given that the system location is 𝑥′ at time 𝑡 = 𝑘∆𝑡 and the component state combination

is at 𝑛′. This system trajectories is typically determined using system code or simulator.

𝑣𝑗′= Volume of the cell 𝑉𝑗′

𝑑𝑥′

𝑣𝑗′= The probability that the state variables are within the infinitesimal volume element 𝑑𝑥′

around 𝑥′ at time 𝑡. This probability is constant throughout each 𝑉𝑗′ (uniform distribution).

Since the functional form of {��(𝑥′, 𝛼𝑛′ , 𝑘∆𝑡)} is generally unknown, the integral in the above

equation is evaluated numerically as follow: [NUREG-6942]

• Partition the cell 𝑗′ into 𝑁𝑝 number of sub-cells

• Choose the midpoint of each sub-cell as initial conditions over the time interval 𝑘𝛥𝑡 ≤

𝑡 ≤ (𝑘 + 1)∆𝑡 under the assumption that the component state combination remains at 𝑛′

at all times during 𝑘𝛥𝑡 ≤ 𝑡 ≤ (𝑘 + 1)∆𝑡

• Observe the number of arrivals in 𝑁𝑝+1 at time 𝑡 = (𝑘 + 1)𝛥𝑡, ��(𝑘+1)∆𝑡(𝑥′, 𝛼𝑛′ , 𝑘∆𝑡)

• Compute the cell-to-cell transition probabilities, 𝑔(𝑗 𝑗′⁄ , 𝑛′, 𝑘∆𝑡) =𝑁𝑝

𝑁𝑝+1

3. Determination of the component state transition probabilities, ℎ(𝑛 𝑛′⁄ , 𝑗′ → 𝑗, 𝑘∆𝑡)

The stochastic behavior of system components is represented by ℎ(𝑛 𝑛′⁄ , 𝑗′ → 𝑗, 𝑘∆𝑡), i.e., the

conditional probability that component state combination is in 𝑛 at time 𝑡 = (𝑘 + 1)𝛥𝑡, given:

Page 73: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

57

• The component state combination was at 𝑛′ at time 𝑡 = 𝑘𝛥𝑡; and

• The state variables make a transition from cell 𝑉𝑗′ to cell 𝑉𝑗 during 𝑘𝛥𝑡 ≤ 𝑡 < (𝑘 + 1)∆𝑡

For system components with statistically dependent failures, the probabilities ℎ(𝑛 𝑛′⁄ , 𝑗′ →

𝑗, 𝑘∆𝑡) is given by the products of the individual component failure or non-failure probabilities

during the mapping time step from 𝑘𝛥𝑡 to (𝑘 + 1)∆𝑡, i.e.,

ℎ(𝑛 𝑛′⁄ , 𝑗′ → 𝑗, 𝑘∆𝑡) = 𝑐𝑚(𝑛𝑚 𝑛′𝑚⁄ , 𝑗′ → 𝑗, 𝑘∆𝑡)

𝑀

𝑚=1

(3-49)

Whereas for statistically independent failure, the probabilities ℎ(𝑛 𝑛′⁄ , 𝑗′ → 𝑗, 𝑘∆𝑡) is given by:

ℎ(𝑛 𝑛′⁄ , 𝑗′ → 𝑗,𝑘∆𝑡) = 𝑐𝑚(𝑛𝑚 𝑛′𝑚⁄ ,𝑘∆𝑡)

𝑀

𝑚=1

(3-50)

Where, 𝑚 = Number of components or units in the system under consideration

𝑐𝑚(𝑛𝑚 𝑛′𝑚⁄ , 𝑗′ → 𝑗, 𝑘∆𝑡): The component state combination at the individual component or unit

level and is the transition probability component 𝑚 from the state combination 𝑛′𝑚 → 𝑛𝑚 during

the time step [𝑘∆𝑡, (𝑘 + 1)∆𝑡].

𝑐𝑚(𝑛𝑚 𝑛′𝑚⁄ , 𝑗′ → 𝑗, 𝑘∆𝑡) can be determined systematically from given component failure rates,

and the set of possible failure modes 𝑆𝑚 for a system transition from cell 𝑉𝑗′ to cell 𝑉𝑗

{𝑖. 𝑒. , 𝑆𝑚(𝑗′, 𝑗)} for each 𝑛𝑚, 𝑛𝑚′ , 𝑗′ and 𝑗.

𝑐𝑚(𝑛𝑚 𝑛′𝑚⁄ , 𝑗′ → 𝑗,∆𝑡) = 1 − 𝜆𝑚 , 𝑛𝑚′

𝑛𝑚′ ∈ 𝑆𝑚 (𝑗 ′ ,𝑗 )

.∆𝑡 → 𝑖𝑓 𝑛𝑚 ∈ 𝑛𝑜𝑟𝑚𝑎𝑙 𝑠𝑡𝑎𝑡𝑒

(3-51)

𝑐𝑚(𝑛𝑚 𝑛′𝑚⁄ , 𝑗′ → 𝑗,∆𝑡) = 𝜆𝑚 , 𝑛𝑚′ ∆𝑡 → 𝑖𝑓 𝑛𝑚 ∈ 𝑓𝑎𝑖𝑙𝑒𝑑 𝑠𝑡𝑎𝑡𝑒

(3-52)

4. Computation of the system transition probability matrix, 𝑞𝑛,𝑗𝑛′,𝑗′

(𝑘∆𝑡) and 𝑃𝑛,𝑗(𝑘∆𝑡)

The stochastic transition probability matrix 𝑞𝑛,𝑗𝑛′,𝑗′

(𝑘∆𝑡) describes the probabilistic dynamic

system behavior and is a function of both ℎ(𝑛 𝑛′⁄ , 𝑗′ → 𝑗, 𝑘∆𝑡) and 𝑔(𝑗 𝑗′⁄ , 𝑛′, 𝑘∆𝑡) with the

Page 74: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

58

matrix having a 𝑁 × 𝐽 dimension. The elements of 𝑞𝑛,𝑗𝑛′,𝑗′(∆𝑡) for the state variable within the

control state space is given by:

𝑞𝑛,𝑗𝑛′,𝑗′(∆𝑡) =

1

𝑣𝑗′ 𝑐𝑚(𝑛𝑚 𝑛′𝑚⁄ , 𝑗′ → 𝑗, ∆𝑡) 𝑑𝑣′𝑒𝑗{��(𝑘+1)∆𝑡(𝑥

′, 𝛼𝑛′ , 𝑘∆𝑡)};𝑗′, 𝑗 = 1,… 𝐽

𝑉𝑗′

𝑀

𝑚=1

Thus, the probability of the system being in state 𝑛 and at cell 𝑗 is simply given by:

𝑃𝑛 ,𝑗 ((𝑘 + 1)∆𝑡) = 𝑞𝑛 ,𝑗𝑛 ′ ,𝑗 ′ (∆𝑡) 𝑃𝑛 ′ ,𝑗 ′ (𝑘∆𝑡)

𝐽

𝑗 ′ =1

𝑁

𝑛 ′ =1

(3-53)

The Markov-CCMT model will be used for its comparison with the classical techniques as well

as with the DFM to meet the objectives of the research. Furthermore, the treatment of transition

probability matrix will be evaluated using a different approach as well as combination of

techniques, such as state merging and transforming the transition matrix to a canonical form.

Page 75: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

59

CHAPTER 4: APPLICATION OF CLASSICAL AND DYNAMIC

METHODOLOGIES

This chapter implements the classical (ET/FT and Markov model) and dynamic PSA (DFM and

Markov-CCMT) methodologies to a benchmark liquid level control system (Aldemir, 1987).

Section 4.1 provide a detail description of the benchmark system (BS), system dynamics and

associated control laws. Section 4.2 and 4.3 presents analysis of the BS using classical ET/FT

analysis. Section 4.4 provide a detail time-dependent analysis (qualitative and quantitative) of the

BS using a classical Markov model. Section 4.5 presents application of DFM to the BS, and

subsequent generation of timed-fault trees from the model. Section 4.6 presents a detail modelling

and analysis of the BS using tightly coupled Markov-CCMT model to generate dynamic accident

sequence. Section 4.7 concludes the chapter with observations made and perform a detail

comparison of classical and dynamic PSA techniques.

The Benchmark System Description

The benchmark example system consists of a fluid hold up tank with three (3) independent fluid

level control units. Each control unit consists of a separate level sensor that measures the fluid

level in the tank, and whose output is an input to the control units. Figure 11 depicts the schematic

diagram of the BS. Operational states of the control units are dependent on the feedback signal

from level sensors. Mission of the BS is to maintain liquid level in the tank to a predefined control

region. The system failure occurs when the liquid level in the tank drained or overflows. Unit-1

and Unit-2 are the two liquid supply units (inflow), whereas Unit-3 is the liquid drain unit

(outflow). Each unit can be thought of as a controller which switches the units “ON” or “OFF”

based on the feedback signal from the level sensors. There are four (4) possible states for each

control units: ON, OFF, fail-open and fail-closed.

The tank liquid level is discretized into an arbitrary number of mutually exclusive regions or

intervals, with the nominal level at 0 meters. All the other possible levels are measured against

the reference level. The maximum level of the tank is 3 meters (at point b) and the minimum level

of the tank is -3 meters (at point a). The system fails if the state variable is outside the defined

control space. Within the maximum and minimum range, two set-points are defined at 𝑎1

(−1meter) and 𝑏1 (+1meter). These set-points partition the control space into three (3) regions

Page 76: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

60

for system operation. The discretization scheme is show in Table 9. Each control region defines

a specific operating states of the units, and hence the system configuration.

Figure 11: The benchmark liquid level control system

Table 9: State variable discretization scheme

Control Region Liquid Level (𝑥) Liquid Level (in meters)

Region 5 𝑥 > 𝑏 Overflow

Region 4 𝑏1 ≤ 𝑥 ≤ 𝑏 +1 ≤ 𝑥 ≤ +3

Region 3 𝑎1 ≤ 𝑥 ≤ 𝑏1 −1 ≤ 𝑥 ≤ +1

Region 2 𝑎 ≤ 𝑥 ≤ 𝑎1 −3 ≤ 𝑥 ≤ −1

Region 1 𝑥 < 𝑎 Drained

Region 1 and Region 5 are considered as the absorbing states in which the top event or system

failure occurs with drained and overflow respectively. No control action is taken once the system

enters the absorbing states. During normal operating condition, the liquid level is in Region 3

(𝑎1 ≤ 𝑥 ≤ 𝑏1). In this region, Unit-1 and Unit-3 are ‘ON’ state balancing the liquid flow-in and

flow-out of the tank, whereas Unit-2 is in standby mode or “OFF” state. When the level transit

Page 77: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

61

from Region 3 to Region 2, Unit-3 receives a signal to turn off whereas Unit-1 and Unit-2 receives

a signal to turn on. Each time the level makes a transition from one region to the other region, the

control system demands certain deterministic actions to all the normal units in order to bring back

the level to the nominal region or at least maintain the system from entering into Region 1 and 5.

The deterministic control laws are given in Table 10.

Table 10: Control laws for the benchmark system

Liquid Level (𝑥) Control Unit State

Unit-1 Unit-2 Unit-3

𝑥 > 𝑏 - - -

𝑏1 < 𝑥 OFF OFF ON

𝑎1 ≤ 𝑥 ≤ 𝑏1 ON OFF ON

𝑥 < 𝑎1 ON ON OFF

𝑥 < 𝑎 - - -

A. Assumptions

The following assumptions are made for the purpose of analysis, which includes:

1. The liquid supply is inexhaustible, and failures of unit are non-repairable;

2. The control units have discrete states and are nominally independent;

3. The response is instantaneous, and the time delay is negligible;

4. The probability of a unit failing-closed and failing-open are assumed to be equally likely

with the failure rates given below (Aldemir, 1987).

𝜆1𝑓𝑐(𝑓𝑎𝑖𝑙𝑒𝑑𝑐𝑙𝑜𝑠𝑒𝑑) = 𝜆1

𝑓𝑜(𝑓𝑎𝑖𝑙𝑒𝑑𝑜𝑝𝑒𝑛) = 2.283 × 10−3/ℎ𝑟.

𝜆2𝑓𝑐(𝑓𝑎𝑖𝑙𝑒𝑑𝑐𝑙𝑜𝑠𝑒𝑑) = 𝜆2

𝑓𝑜(𝑓𝑎𝑖𝑙𝑒𝑑𝑜𝑝𝑒𝑛) = 2.857 × 10−3/ℎ𝑟

𝜆3𝑓𝑐(𝑓𝑎𝑖𝑙𝑒𝑑𝑐𝑙𝑜𝑠𝑒𝑑) = 𝜆3

𝑓𝑜(𝑓𝑎𝑖𝑙𝑒𝑑𝑜𝑝𝑒𝑛) = 1.563 × 10−3/ℎ𝑟

B. Benchmark System Dynamics

The tank is considered to be at the nominal level (Region 3) at the start of the system operation.

If any transient occurs due to failure of a unit, the state variable can transit out of the nominal

control region which can lead to system failure or can be kept within control space depending on

the operational states of the remaining units. The state variable transition to failure space is

Page 78: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

62

dependent on the initial level as well as the rate of change of liquid level in the tank. Since the

initial liquid level is a known input, the element of interest is the rate of liquid level change. The

system dynamics is given by:

𝑑𝑥(𝑡)

𝑑𝑡= 𝑓𝑛(𝑥)

(4-1)

Where; 𝑥 = The liquid level in the tank

𝑓𝑛 = Rate of change of liquid level as a function of control unit states (𝑛).

Table 11 list 𝑓𝑛 with respect to the unit states, where ��1, ��2 and ��3 are the flowrates from Unit-1,

Unit-2 and Unit-3 respectively. The flowrates at states fail-open and fail-closed for the units are

not depicted in the table since the rate of change of level are the same with units’ states ON and

OFF respectively. Considering an infinitesimally small time step (∆𝑡) relative to the system

dynamics and failure rate of the units, and with the assumption that the system configuration does

not change during this small interval, the liquid level in the tank at (𝑡 + ∆𝑡) can be represented

by the following equation.

𝑥(𝑡 + ∆𝑡) ≈ 𝑥′(𝑡) + 𝑓𝑛 ′ (𝑥′).∆𝑡

(4-2)

Where; 𝑥′(𝑡) = Liquid level at time 𝑡

𝑓𝑛′(𝑥′) = Rate of change of liquid level as a function of system configuration 𝑛′

𝑥(𝑡 + ∆𝑡) = Liquid level at time (𝑡 + ∆𝑡)

Table 11: Total net fluid flow into/out of the tank

Control Unit State Rate of Level

change (𝑓𝑛)

Net fluid flow

Unit-1 Unit-2 Unit-3

ON ON ON ��1 + ��2 − ��3 ��

ON ON OFF ��1 + ��2 2��

ON OFF ON ��1 − ��3 0

ON OFF OFF ��1 ��

OFF ON ON ��2 − ��3 0

OFF ON OFF ��2 ��

Page 79: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

63

OFF OFF ON −��3 −��

OFF OFF OFF 0 0

As mentioned earlier, the system is assumed to be in the nominal control region, where Unit-1 is

ON, Unit-2 is OFF and Unit-3 is ON. This operating condition will continue until one of the units

make a transition from normal to failure states. This transition to failure state will cause the fluid

level to deviate from its nominal value, which can be level high or level low depending on which

unit has failed and to which failure state. The system will change its configuration based on the

new control region with the objective to bring back the level to its nominal value. The new system

state may or may not result to a stable condition.

Fault Tree Analysis of the Benchmark System

Fault tree analysis for the benchmark system can be performed in two different ways depending

on the definition of the top events. Fault trees for:

1. Binary unit/system state (normal or failure state);

2. Multiple unit/system states (normal, failed-closed and failed-open states).

Fault tree following the first approach (binary states) can be constructed as shown in Figure 12.

The top event probability is evaluated as:

Top Event = 𝐺1𝐺2 + 𝐺3 = 3.15 × 10−3

Figure 12: Fault tree for the benchmark system failure (binary state)

Page 80: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

64

The second approach is necessary when top events are defined in terms of system state or state

variable magnitudes. Since the benchmark system failure is defined in terms of state variable

magnitude, it is a more natural way to construct FTs using the second approach. FTs for the

benchmark system overflow and drained as shown in Figure 13 and Figure 14 respectively.

Figure 13: Fault tree for the benchmark system drained

Minimal cut-set for BS drained = 𝐵1𝐵2𝐵3 = 1.02𝐸 − 08

Minimal cut-set for BS overflow = 𝐵1𝐵2 + 𝐵1𝐵3 + 𝐵2𝐵3 = 1.455𝐸 − 05

Page 81: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

65

Figure 14: Fault Tree for the Benchmark System Overflow

Page 82: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

66

Event Tree Analysis of the Benchmark System

Provided an initiating event (IE), the state ordering or sequential operation of the frontline

systems/units is fixed for a classical ET. The unit response, accident sequence and hence the final

system state is significantly affected by the modes of unit failure. Furthermore, the sequence of

unit state transition can affect the type of system failure or may lead the system to a quasi-stable

state. Whereas in case of a FT, all sequence or cut-sets result to a system failure which is not

necessarily true in this case. Since the benchmark system responds dynamically to any IEs, an ET

is a more essential approach than FT for modeling the benchmark system. The unit failure modes

have been explicitly modelled in the ET since different modes lead to different system states. For

analysis purpose, hardware-oriented ETs are constructed for the system's response to the following

IEs. These IEs was selected since they possess an interesting case and reveals some limitations of

the technique.

1. Unit-1 failed-closed and failed-open;

2. Unit-2 failed-open;

3. Unit-3 failed-closed.

For the ETs constructed, quantification of the accident sequence is straight forward as can be

observed in Figure 15. The analysis was performed using CAFTA code. Only the first case is

illustrated for discussion purpose.

a. Unit-1 Failed-Closed (𝝀𝟏𝒇𝒄

)

Event tree with an IE probability of 𝜆1𝑓𝑐= 2.28 × 10−3/ℎ𝑟 is depicted in Figure 15. It can be

observed that with an IE Unit-1 failed-closed, probability of the benchmark system failing by

overflow and drained condition is equally likely with a probability of 1.02E − 08/ℎ𝑟. Hence, there

is a 50% probability that the system will overflow or drained if unit-1 failed-closed. The argument

for this can be made as; given that unit-1 failed closed, liquid level in the tank will decrease with

time since unit-3 is in normal state. However, when the liquid level transit below the nominal

region, unit-2 will be turned on and unit-3 will be turned off as defined in the control laws. This

action of unit-2 and unit-3 will cause the liquid level to rise till it reaches the nominal region. Once

the level is in the nominal region, unit-2 will be tuned off and hence the level will start to decrease

again. This oscillation will take place until one of the two units failure occur. Given that the unit-

Page 83: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

67

3 and unit-2 failing closed and failing open respectively have the same probabilities, it can be

concluded that there is a 50% probability for system overflow and drained scenario. Also, it is

evident that the system will be operating mostly around liquid level 𝑎1. Hence, after a qualitative

analysis of the evolution of state variable, one may conclude that it is more likely for the system

to fail by drained than overflow, or at least it will take more amount of time of the system to

overflow for this particular scenario. However, this conclusion can only be definitive and

quantified when the analyst has information about the timing of failure event.

Note that the ordering of the unit response to the IE have been fixed a prior by the analyst, i.e.,

unit-3 response first to the IE and then the unit-2. From the figure, it can be observed that Unit-3

being normal will lead the system to a stable or quasi stable state. Unit-3 failing open and unit-2

failing close will lead the system to a drained scenario. Similarly, unit-3 failing-closed and unit-2

failing open will result to a system overflow condition. It is important to note here that the failure

of two (2) or three (3) units does not necessarily result to a system failure condition. For instance,

the sequence Unit-1 failed-closed AND Unit-3 failed-open AND Unit-1 failed-open lead to a

system quasi-stable scenario, i.e., the liquid level in the nominal region even after a failure of all

the units. In fact, only two (2) event sequence lead to a system failure condition out of nine (9)

sequences generated.

Figure 15: Event tree for initiating event “Unit-1 Failed-Closed”

The probability of the benchmark system overflow or drained scenario can be calculated by

summing up all the accident sequences leading to a system overflow or drained respectively:

Page 84: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

68

System overflow = 2.91𝐸 − 05

System drained = 3.06𝐸 − 08

It is evident that ET approach provide the analyst with more information about the dynamic system

response and the possible end states compared to the FT technique. However, as mentioned earlier,

the event ordering is fixed which may not be true in some scenarios. Again, the system response

is dependent on evolution of the state variables when the next failure occurs in the sequence, timing

of the failure event and pre-defined deterministic control laws. For example, Unit-2 may respond

first in contrast to Unit-3 for IE unit-1 failed-closed. The ETs presented above explicitly models

the different failure modes of the units, since the change in failure modes can result to a different

system states or type. Thus, ETs must be constructed for each IEs, which becomes tedious for a

system with several IEs. However, this approach enables the analyst to observe all possible

scenarios for an IE, which can further be treated using well-established system codes. For instance,

an accident sequence may lead the BS to quasi stable or failure state; however, the time taken from

the system to reach these states, and final liquid level in the tank can be determined only if the

system dynamics is considered.

Markov Model of the Benchmark System

This section presents the time-dependent stochastic modelling of the BS using discrete-state

discrete-time Markov chain. The time dependent behavior of the BS can be modelled by

constructing state transition diagrams consisting of system states and possible transitions among

these states. State ordering is explicitly modeled in defining the possible transitions in/out of a

system states. Thus, Markov model provide a possible solution to overcome the limitations of

capturing the time element and state ordering which was encountered in FT/ET techniques.

Furthermore, the methodology provides a superior approach and solution for modeling

components/units with multiple failure modes and system states. In constructing the Markov

transition diagram, some assumptions were made, including:

1. The units are non-repairable, and the units fail only once in a given failure sequence, i.e.,

once a unit failure mode have occurred, other modes cannot occur for the same unit. This

assumption is true since units failing closed/open have different effect on the system and

must be modelled separately;

Page 85: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

69

2. The system is initially at normal operating condition, i.e., all units are normal and liquid

level is at the nominal region at 𝑡 = 0;

3. The system states are defined uniquely by distinct combination of the individual unit states

and are statistically independent.

At this point it should be made clear that for a unit in normal state, the unit can either be in ON or

OFF depending on the user defined control laws, which in turn is characterized by the liquid level

in the tank. Hence, the ON and OFF states can be combined and represented by a normal state, as

was done while constructing the transition diagram (See Figure 16). Since each of the units have

three (3) distinct states, the total number of distinct states that the system can have is:

𝑁 = (𝑃𝑜𝑠𝑠𝑖𝑏𝑙𝑒𝑓𝑎𝑖𝑙𝑢𝑟𝑒𝑚𝑜𝑑𝑒𝑠)(𝑛𝑢𝑚𝑏𝑒𝑟𝑜𝑓𝑢𝑛𝑖𝑡𝑠) = 33

𝑁 = 27 system states

The 27 distinct system states can be written as an ordered set of distinct unit state or sub-sets for

ease of system modelling with Markov chain. Consider, 𝑛1 = 0 implies Unit-1 in normal state;

𝑛1 = 1 implies Unit-1 in failed-closed state; and 𝑛1 = 2 implies Unit-1 in failed-open state.

Similarly, the same is true for Unit-2 and Unit-3. The possible unit state combinations which was

used to construct the transition diagram is presented in Table 12.

The transition diagram starts with all the units being in normal state at 𝑡 = 0, i.e., system state 𝑛 =

0. All the possible transition out of the normal system state are the sum of all the individual failure

modes of each unit, i.e., system states with a single failure. Similarly, as the time step increase the

transition from a single unit failure to two (2) unit failure can occur, and subsequently from a two

(2) unit failure to three (3) unit failure, i.e., each additional failure generates a new system state.

Hence, the transition diagram can be represented in a layer wise, with four (4) possible transition

layers in the Markov model (See Table 13).

Page 86: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

70

Table 12: Possible unit state combination

Individual Unit States System States

(𝑛)

Possible end states

Unit-1 (𝑛1) Unit-2 (𝑛2) Unit-3 (𝑛3)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 0 0 𝑃0(𝑡) System stable

𝑛1 = 1 𝑛2 = 0 𝑛3 = 0 1 𝑃1(𝑡) System stable

𝑛1 = 2 𝑛2 = 0 𝑛3 = 0 2 𝑃2(𝑡) System stable

𝑛1 = 0 𝑛2 = 1 𝑛3 = 0 3 𝑃3(𝑡) System stable

𝑛1 = 0 𝑛2 = 2 𝑛3 = 0 4 𝑃4(𝑡) System stable

𝑛1 = 0 𝑛2 = 0 𝑛3 = 1 5 𝑃5(𝑡) System stable

𝑛1 = 0 𝑛2 = 0 𝑛3 = 2 6 𝑃6(𝑡) System stable

𝑛1 = 1 𝑛2 = 1 𝑛3 = 0 7 𝑃7(𝑡) System quasi-stable

𝑛1 = 1 𝑛2 = 2 𝑛3 = 0 8 𝑃8(𝑡) System quasi-stable

𝑛1 = 1 𝑛2 = 0 𝑛3 = 1 9 𝑃9(𝑡) System quasi-stable

𝑛1 = 1 𝑛2 = 0 𝑛3 = 2 10 𝑃10(𝑡) System quasi-stable

𝑛1 = 2 𝑛2 = 1 𝑛3 = 0 11 𝑃11(𝑡) System quasi-stable

𝑛1 = 2 𝑛2 = 2 𝑛3 = 0 12 𝑃12(𝑡) System Overflow

𝑛1 = 2 𝑛2 = 0 𝑛3 = 1 13 𝑃13(𝑡) System Overflow

𝑛1 = 2 𝑛2 = 0 𝑛3 = 2 14 𝑃14(𝑡) System quasi-stable

𝑛1 = 0 𝑛2 = 1 𝑛3 = 1 15 𝑃15(𝑡) System quasi-stable

𝑛1 = 0 𝑛2 = 1 𝑛3 = 2 16 𝑃16(𝑡) System quasi-stable

𝑛1 = 0 𝑛2 = 2 𝑛3 = 1 17 𝑃17(𝑡) System Overflow

𝑛1 = 0 𝑛2 = 2 𝑛3 = 2 18 𝑃18(𝑡) System quasi-stable

𝑛1 = 1 𝑛2 = 1 𝑛3 = 1 19 𝑃19(𝑡) System quasi-stable

𝑛1 = 1 𝑛2 = 1 𝑛3 = 2 20 𝑃20(𝑡) System Dryout

𝑛1 = 1 𝑛2 = 2 𝑛3 = 1 21 𝑃21(𝑡) System Overflow

𝑛1 = 1 𝑛2 = 2 𝑛3 = 2 22 𝑃22(𝑡) System quasi-stable

𝑛1 = 2 𝑛2 = 1 𝑛3 = 1 23 𝑃23(𝑡) System Overflow

𝑛1 = 2 𝑛2 = 1 𝑛3 = 2 24 𝑃24(𝑡) System quasi-stable

𝑛1 = 2 𝑛2 = 2 𝑛3 = 1 25 𝑃25(𝑡) System Overflow

𝑛1 = 2 𝑛2 = 2 𝑛3 = 2 26 𝑃26(𝑡) System Overflow

Page 87: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

71

The system states are depicted systematically in a sequential layer representation. The first possible

transition from normal system state is presented in Figure 16. Note that the order of failures should

be strictly followed when generating failure sequences from the model because the variation in

state ordering can produce different end states. Since the number of system states and the number

of possible transition from one state to the other state is large, it is easier to construct the transition

diagram considering a specific failure mode. Figure 17 depict the possible transition path from

Layer-1 to Layer-3 (only one path depicted for illustration purpose).

Table 13: Transitional layers representation of system states

No. Layers System State

1. Layer-0 All units in normal state

2. Layer-1 One (1) unit failure

3. Layer-2 Two (2) units failure

4. Layer-3 Three (3) units failure

Figure 16: Transition diagram from Layer-1 to Layer-2

Layer-0 Layer-1

Page 88: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

72

Figure 17: Transition path from Layer-0 to Unit-1 failed-closed to Layer-3

From the constructed Markov transition diagram, a set of ODEs can be obtained. 1

𝑑𝑃0(𝑡)

𝑑𝑡= −(𝜆1

𝑓𝑐+ 𝜆1

𝑓𝑜+ 𝜆2

𝑓𝑐+ 𝜆2

𝑓𝑜+ 𝜆3

𝑓𝑐+ 𝜆3

𝑓𝑜)𝑃0(𝑡)

𝑑𝑃1(𝑡)

𝑑𝑡= 𝜆1

𝑓𝑐𝑃0(𝑡) − (𝜆2

𝑓𝑐+ 𝜆2

𝑓𝑜+ 𝜆3

𝑓𝑐+ 𝜆3

𝑓𝑜)𝑃1(𝑡)

𝑑𝑃2(𝑡)

𝑑𝑡= 𝜆1

𝑓𝑜𝑃0(𝑡) − (𝜆2

𝑓𝑐+ 𝜆2

𝑓𝑜+ 𝜆3

𝑓𝑐+ 𝜆3

𝑓𝑜)𝑃2(𝑡)

𝑑𝑃3(𝑡)

𝑑𝑡= 𝜆2

𝑓𝑐𝑃0(𝑡) − (𝜆1

𝑓𝑐+ 𝜆1

𝑓𝑜+ 𝜆3

𝑓𝑐+ 𝜆3

𝑓𝑜)𝑃3(𝑡)

𝑑𝑃4(𝑡)

𝑑𝑡= 𝜆2

𝑓𝑜𝑃0(𝑡) − (𝜆1

𝑓𝑐+ 𝜆1

𝑓𝑜+ 𝜆3

𝑓𝑐+ 𝜆3

𝑓𝑜)𝑃4(𝑡)

𝑑𝑃5(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑐𝑃0(𝑡) − (𝜆1

𝑓𝑐+ 𝜆1

𝑓𝑜+ 𝜆2

𝑓𝑐+ 𝜆2

𝑓𝑜)𝑃5(𝑡)

𝑑𝑃6(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑜𝑃0(𝑡) − (𝜆1

𝑓𝑐+ 𝜆1

𝑓𝑜+ 𝜆2

𝑓𝑐+ 𝜆2

𝑓𝑜)𝑃6(𝑡)

𝑑𝑃7(𝑡)

𝑑𝑡= 𝜆2

𝑓𝑐𝑃1(𝑡) + 𝜆1

𝑓𝑐𝑃3(𝑡) − (𝜆3

𝑓𝑐+ 𝜆3

𝑓𝑜)𝑃7(𝑡)

𝑑𝑃8(𝑡)

𝑑𝑡= 𝜆2

𝑓𝑜𝑃1(𝑡) + 𝜆1

𝑓𝑐𝑃4(𝑡) − (𝜆3

𝑓𝑐+ 𝜆3

𝑓𝑜)𝑃8(𝑡)

1 Note that all the 27 ODEs are grouped and depicted by a single equation numbering scheme for easier representation

Layer-2 Layer-1 Layer-3 Layer-0

Page 89: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

73

𝑑𝑃9(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑐𝑃1(𝑡) + 𝜆1

𝑓𝑐𝑃5(𝑡) − (𝜆2

𝑓𝑐+ 𝜆2

𝑓𝑜)𝑃9(𝑡)

𝑑𝑃10(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑜𝑃1(𝑡) + 𝜆1

𝑓𝑐𝑃5(𝑡) − (𝜆2

𝑓𝑐+ 𝜆2

𝑓𝑜)𝑃10(𝑡)

𝑑𝑃11(𝑡)

𝑑𝑡= 𝜆2

𝑓𝑐𝑃2(𝑡) + 𝜆1

𝑓𝑜𝑃3(𝑡) − (𝜆3

𝑓𝑐+ 𝜆3

𝑓𝑜)𝑃11(𝑡)

𝑑𝑃12(𝑡)

𝑑𝑡= 𝜆2

𝑓𝑜𝑃2(𝑡) + 𝜆1

𝑓𝑜𝑃4(𝑡) − (𝜆3

𝑓𝑐+ 𝜆3

𝑓𝑜)𝑃12(𝑡)

𝑑𝑃13(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑐𝑃2(𝑡) + 𝜆1

𝑓𝑜𝑃5(𝑡) − (𝜆2

𝑓𝑐+ 𝜆2

𝑓𝑜)𝑃13(𝑡)

𝑑𝑃14(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑜𝑃2(𝑡) + 𝜆1

𝑓𝑜𝑃6(𝑡) − (𝜆2

𝑓𝑐+ 𝜆2

𝑓𝑜)𝑃14(𝑡)

𝑑𝑃15(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑐𝑃3(𝑡) + 𝜆2

𝑓𝑐𝑃5(𝑡) − (𝜆1

𝑓𝑐+ 𝜆1

𝑓𝑜)𝑃15(𝑡)

𝑑𝑃16(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑜𝑃3(𝑡) + 𝜆2

𝑓𝑐𝑃6(𝑡) − (𝜆1

𝑓𝑐+ 𝜆1

𝑓𝑜)𝑃16(𝑡)

𝑑𝑃17(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑐𝑃4(𝑡) + 𝜆2

𝑓𝑜𝑃5(𝑡) − (𝜆1

𝑓𝑐+ 𝜆1

𝑓𝑜)𝑃17(𝑡)

𝑑𝑃18(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑜𝑃4(𝑡) + 𝜆2

𝑓𝑜𝑃6(𝑡) − (𝜆1

𝑓𝑐+ 𝜆1

𝑓𝑜)𝑃18(𝑡)

𝑑𝑃19(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑐𝑃7(𝑡) + 𝜆2

𝑓𝑐𝑃9(𝑡) + 𝜆1

𝑓𝑐𝑃15(𝑡)

𝑑𝑃20(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑜𝑃7(𝑡) + 𝜆2

𝑓𝑐𝑃10(𝑡) + 𝜆1

𝑓𝑐𝑃16(𝑡)

𝑑𝑃21(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑐𝑃8(𝑡) + 𝜆2

𝑓𝑜𝑃9(𝑡) + 𝜆1

𝑓𝑐𝑃17(𝑡)

𝑑𝑃22(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑜𝑃8(𝑡) + 𝜆2

𝑓𝑜𝑃10(𝑡) + 𝜆1

𝑓𝑐𝑃18(𝑡)

𝑑𝑃23(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑐𝑃11(𝑡) + 𝜆2

𝑓𝑐𝑃13(𝑡) + 𝜆1

𝑓𝑜𝑃15(𝑡)

𝑑𝑃24(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑜𝑃11(𝑡) + 𝜆2

𝑓𝑐𝑃14(𝑡) + 𝜆1

𝑓𝑜𝑃16(𝑡)

𝑑𝑃25(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑐𝑃12(𝑡) + 𝜆2

𝑓𝑜𝑃13(𝑡) + 𝜆1

𝑓𝑜𝑃17(𝑡)

𝑑𝑃26(𝑡)

𝑑𝑡= 𝜆3

𝑓𝑜𝑃12(𝑡) + 𝜆2

𝑓𝑜𝑃14(𝑡) + 𝜆1

𝑓𝑜𝑃18(𝑡)

(4-3)

The above ODEs can be solved with given initial conditions. However, before going ahead to

solving the above transition rate equations, a qualitative overview is provided for better

comprehension of the problem at hand.

Page 90: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

74

4.4.1. Qualitative Assessment

A preliminary qualitative analysis can be performed with the knowledge of the system states

defined in the previous Table 12. It is important to note here that this qualitative assessment is only

an approximation. It can be observed that failure probability of the BS to maintain nominal liquid

level with all possible IEs is 29.63%. Whereas, probability of the BS failing by overflow and

drained is 87.5% and 12.5% respectively.

Taking a step further, if the order and sequence of unit failure is strictly taken into account, the

probability of a unit failing first must be considered and treated separately. The probability that a

unit will fail first is given by the ratio of failure rate for that unit divided by the sum of the failure

rates for all the units. Thus, probability of Unit-1 failing first is:

𝑃(Unit − 1 → failingfirst) =𝜆1

𝜆1+𝜆2+𝜆3 (4-4)

𝑃(Unit − 1 → failingfirst) = 34.06%

Similarly, the probability of Unit-2 and Unit-3 failing first is determined to be:

𝑃(Unit − 2 → failingfirst) = 42.63%

𝑃(Unit − 3 → failingfirst) = 23.31%

The strict systematic ordering and sequencing of transition, and its subsequent end state with all

possible unit failure modes is shown in Figure 18. This approach has some differences with the

conventional Markov chain, in the sense that the number of end states and absorbing states are

different.

It can be observed that all system end state with failure modes ‘Unit-1 failed-open’ will eventually

lead to a system overflow scenario or an absorbing state. Hence, with the failure mode ‘Unit-1

failed-open’, there is a 100% probability that the system will make a transition to an absorbing

state or more specifically, the system will fail by overflow condition. For Unit-1 failed-closed

failure mode, the liquid level will decrease until it reaches Region -1. Then, Unit-2 will be turned

on and Unit-3 will be turned off as per control laws defined with respect to the liquid level. This

causes the fluid level to rise until it reaches Region +1, at which time Unit-2 will be turned off and

Unit-3 will be tuned on. This causes the liquid level to uniformly oscillate between 𝑅𝑒𝑔𝑖𝑜𝑛 − 1

and 𝑅𝑒𝑔𝑖𝑜𝑛 + 1. Since the liquid flowrates from all the units are equal, the BS spends an equal

Page 91: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

75

amount of time in the increasing and decreasing liquid level conditions. Thus, it can be concluded

that, the BS is equally likely to fail in an overflow or drained state. Therefore, with the failure

mode Unit-1 failed-closed, there is a 50% chance that the system will fail by overflow and drained

respectively. Combining all the possible system failures for both the failure modes from Unit-1

(failed-closed + failed-open), it can be calculated that the probability of BS failing by overflow

and drained is 80% and 20% respectively. Similarly, the state sequences transition with failure

modes Unit-2 failed-closed and failed-open are the same as Unit-1.

Figure 18: System state transition for Unit-1 failed-open

For Unit-3 (the only drain unit), all system failures caused by the failure mode Unit-3 failed-open,

there is a 50% probability of system failing by overflow and drained respectively. This is true

since the flowrates from all the units are assumed to be equal. For the failure mode Unit-3 failed-

closed, there is 100% probability that the system will fail by overflow condition regardless of the

relative flow rates provided by the three units. Unit-3 failing-closed will cause the liquid level to

rise until it crosses the nominal region, at which time Unit-1 will be turned off. The tank will

remain in this condition until either Unit-1 or Unit-2 fail-open, either of which will lead directly

to the BS overflow. Thus, the probability of system failing by overflow and drained is 80% and

Page 92: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

76

20% respectively. Utilizing the calculated probability values for individual units failing first, the

probability that the BS failing by overflow is determined as:

(0.3406213 × 0.8) + (0.4262653 × 0.8) + (0.233113 × 0.8) = 0.80

Thus, the BS will fail by overflow condition approximately 80% of the time and fail by system

drained approximately 20% of the time. Note that, the system being stable or quasi-stable are not

considered for the current analysis since we are interested only on the state ordering and qualitative

aspect of the sequence transition and absorbing states. These aspects will be treated later when

Markov model is coupled with CCMT to capture the system dynamics.

The strict states ordering results in 48 total numbers of failure sequences/end states in contrast to

the 27 system states generated by applying the classical Markov chain. The failure probability of

the system states generated by the two approaches are different, wherein the first approach

overestimates the probability of system failing by overflow and underestimates the probability of

system failing by drained. The differences in the two approaches are presented in Table 14.

Considering that all unit flowrates are equal for the BS, and that some unit failure type does not

necessarily lead to a change in system configuration due to the control laws, these unit failure type

can be omitted from IEs. For instance, if the system is in normal operating condition, Unit-3 failing

open does not lead to a change in system configuration since Unit-1 fully open can compensate for

these failures. The transition probabilities without a system configuration change is zero, in which

case, the IE can be omitted. Similarly, the same argument holds for Unit-1 failing-open and Unit-

2 failing-closed since these failures does not lead to a change in system configuration. Note that

this consideration is true only if the system is under normal operating condition. Thus, considering

only the three (3) IEs; namely Unit-1 failed-closed, Unit-2 failed-open and Unit-3 failed-closed,

the qualitative assessment can be performed for the system evolution leading to 18 failure

sequences. Taking the IE Unit-3 failed-closed, all system failure event sequence eventually leads

to system overflow condition i.e., a 100% failure probability. For Unit-1 failed-closed, it is similar

to as described in the previous section, i.e., there is a 50% chance that the system will fail by

overflow and drained respectively. Similarly, for the IE Unit-2 failed-open, there is a 100 %

probability that the system will fail by overflow condition. Utilizing the calculated probability for

individual units failing first, the probability that the BS failing by overflow is:

Page 93: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

77

(0.341 × 50%) + (0.426 × 100%) + (0.233 × 100%) = 0.83

Thus, the BS will fail by overflow condition approximately 83% of the time and fail by system

drained approximately 17% of the time. Differences in the qualitative assessment are presented

in Table 14

Table 14: A qualitative comparison of three different approach

Approaches Number of Failure

Sequence

System failure states

Overflow (%) Drained (%)

Conventional Markov chain 27 87.50% 12.50%

Markov chain with explicit

state ordering 48 80.00% 20.00%

Markov chain with explicit

state ordering for 3 IEs 18 83.00% 17.00%

The third approach provide a good approximation in the sense that the system failures by overflow

and drained is not overestimated or under estimated. This is due to the fact that a specific unit

failure is not necessarily an IE leading to a system failure. Hence, these failures are not included

as IEs for analysis purpose. Of course, like all Markov chain state representation, this

approximation is only true with the assumption that two failures do not occur simultaneously in a

single time step.

4.4.2. Quantitative Assessment

The ordinary differential equations (ODEs) obtained from the Markov state transition diagram can

be solved using two approaches: (1) Analytical methods; (2) Numerical methods.

1. Analytical methods

From the assumption that the BS is in normal operating condition at time 𝑡 = 0, or that the system

is in state 𝑃0(𝑡 = 0), the initial conditions for all the ODEs can be defined as:

𝑃0(0) = 1;𝑃1(0) = 0;𝑃2(0) = 0;𝑃3(0) = 0;𝑃4(0) = 0;……𝑃26(0) = 0

The solution of the transition probabilities with given initial conditions are:

𝑃0(𝑡) = exp{−(𝜆1𝑓𝑐+ 𝜆1

𝑓𝑜+ 𝜆2

𝑓𝑐+ 𝜆2

𝑓𝑜+ 𝜆3

𝑓𝑐+ 𝜆3

𝑓𝑜)𝑡} (4-5)

Page 94: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

78

𝑃7(𝑡) =𝜆1𝑓𝑐𝜆2𝑓𝑐

𝜆1𝑓𝑐𝜆2𝑓𝑐

+ 𝜆1𝑓𝑐𝜆2𝑓𝑜

+ 𝜆1𝑓𝑜𝜆2𝑓𝑐

+ 𝜆1𝑓𝑜𝜆2𝑓𝑜𝑒− 𝜆1

𝑓𝑐+𝜆1

𝑓𝑜+𝜆2

𝑓𝑐+𝜆2

𝑓𝑜+𝜆3

𝑓𝑐+𝜆3

𝑓𝑜 𝑡

−𝜆1𝑓𝑐𝜆2𝑓𝑐

𝜆1𝑓𝑐𝜆2𝑓𝑐

+ 𝜆1𝑓𝑐𝜆2𝑓𝑜

+ 𝜆1𝑓𝑜𝜆2𝑓𝑐

+ 𝜆1𝑓𝑜𝜆2𝑓𝑜𝑒− 𝜆1

𝑓𝑐+𝜆1

𝑓𝑜+𝜆3

𝑓𝑐+𝜆3

𝑓𝑜 𝑡

−𝜆1𝑓𝑐𝜆2𝑓𝑐

𝜆1𝑓𝑐𝜆2𝑓𝑐

+ 𝜆1𝑓𝑐𝜆2𝑓𝑜

+ 𝜆1𝑓𝑜𝜆2𝑓𝑐

+ 𝜆1𝑓𝑜𝜆2𝑓𝑜𝑒− 𝜆2

𝑓𝑐+𝜆2

𝑓𝑜+𝜆3

𝑓𝑐+𝜆3

𝑓𝑜 𝑡

+𝜆1𝑓𝑐𝜆2𝑓𝑐

𝜆1𝑓𝑐𝜆2𝑓𝑐

+ 𝜆1𝑓𝑐𝜆2𝑓𝑜

+ 𝜆1𝑓𝑜𝜆2𝑓𝑐

+ 𝜆1𝑓𝑜𝜆2𝑓𝑜𝑒− 𝜆3

𝑓𝑐+𝜆3

𝑓𝑜 𝑡

(4-6)

And so on.

Solving the coupled ODEs via analytical methods become extremely complex and time

consuming. Furthermore, with an increase in the number of component states, it becomes

impractical to solve ODEs by analytical methods. Thus, numerical methods provide a better

approach in solving the coupled ODEs.

2. Numerical methods

Numerical method was used in this thesis for obtaining the solution of the coupled ODEs since it

provides a more robust and practical approach. Finite difference method (FDM) has been used for

solving the coupled ODEs. The transition probability ODEs can be written in state matrix form as:

𝑑𝑃(𝑡)

𝑑𝑡= 𝐴.𝑃(𝑡)

(4-7)

Where, the matrix “A” is the transition probability matrix, and can be validated by using:

𝐴𝑖𝑗

𝑛

𝑖=1

𝑛

𝑗=1

= 1

(4-8)

Left hand side of Equation (35) can be re-written in the following way;

𝑑𝑃(𝑡)

𝑑𝑡=𝑃(𝑡 + ∆𝑡) − 𝑃(𝑡)

(𝑡 + ∆𝑡) − 𝑡

(4-9)

Substituting Equation (4-5) in Equation (4-3) yields:

Page 95: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

79

𝑃(𝑡 + ∆𝑡) − 𝑃(𝑡)

∆𝑡= 𝐴.𝑃(𝑡)

𝑃(𝑡 + ∆𝑡) = (𝐼 + 𝐴.∆𝑡).𝑃(𝑡)

(4-10)

The above equation is equivalent to the Chapman–Kolmogorov equation in discrete time state-

space. With the assumption that the BS is in normal operating condition at time 𝑡 = 0, the state

transition probabilities are computed by implementing FDM in FORTRAN95 code. Considering

a time step of ∆𝑡 = 1ℎ𝑜𝑢𝑟, we obtain the transition probabilities for 3 time steps (See Table 15).

At any given time step, system failure probability can be determined by summing up the states

leading to either system overflow or drained scenario. For instance, from Table 15, it can be

observed that states 12, 13, 17, 21, 23, 25 and 26 result in a system overflow scenario. For

illustration purpose, the probability of system overflow at time step 𝑘 = 4 is determined as:

System overflow (𝑘 = 4) = 𝑃12(𝑡) + 𝑃13(𝑡) + 𝑃17(𝑡) + 𝑃21(𝑡) + 𝑃23(𝑡) + 𝑃25(𝑡) + 𝑃26(𝑡)

System overflow (𝑘 = 4) = 8.68 × 10−05

Similarly, the probability of system drained at 𝑘 = 4 can be obtained as:

System drained (𝑘 = 4) = 𝑃20(𝑡)

System drained (𝑘 = 4) = 6.12 × 10−08

Similarly, the probability of system failure (overflow or drained) for any given time step can be

obtained using the same procedure as depicted above.

Page 96: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

80

Table 15: Transition probabilities in terms of 𝑃𝑛(𝑡)

# 𝑃0(𝑡) 𝑃1(𝑡) 𝑃2(𝑡) 𝑃3(𝑡) 𝑃4(𝑡) 𝑃5(𝑡) 𝑃6(𝑡) 𝑃7(𝑡) 𝑃8(𝑡)

𝑘 = 1 1 0 0 0 0 0 0 0 0

𝑘 = 2 0.987 2.28E-3 2.28E-3 2.86E-3 2.857E-3 1.563E-3 1.563E-3 0 0

𝑘 = 3 0.974 4.52E-3 4.52E-3 5.65E-03 5.654E-03 3.088E-03 3.088E-03 1.305E-5 1.305E-5

𝑘 = 4 0.96 6.698E-3 6.698E-3 8.39E-3 8.392E-3 4.577E-03 4.577E-03 3.88E-5 3.882E-5

# 𝑃9(𝑡) 𝑃10(𝑡) 𝑃11(𝑡) 𝑃12(𝑡) 𝑃13(𝑡) 𝑃14(𝑡) 𝑃15(𝑡) 𝑃16(𝑡) 𝑃17(𝑡)

𝑘 = 1 0 0 0 0 0 0 0 0 0

𝑘 = 2 0 0 0 0 0 0 0 0 0

𝑘 = 3 7.12E-6 7.135E-6 1.305E-5 1.30E-5 7.135E-6 7.13E-6 8.929E-6 8.93E-6 8.929E-6

𝑘 = 4 2.11E-05 2.11E-05 3.882E-5 3.88E-5 2.12E-05 2.12E-05 2.66E-05 2.66E-05 2.66E-05

# 𝑃18(𝑡) 𝑃19(𝑡) 𝑃20(𝑡) 𝑃21(𝑡) 𝑃22(𝑡) 𝑃23(𝑡) 𝑃24(𝑡) 𝑃25(𝑡) 𝑃26(𝑡)

𝑘 = 1 0 0 0 0 0 0 0 0 0

𝑘 = 2 0 0 0 0 0 0 0 0 0

𝑘 = 3 8.93E-6 0 0 0 0 0 0 0 0

𝑘 = 4 2.65E-05 6.12E-08 6.12E-08 6.12E-08 6.12E-08 6.12E-08 6.12E-08 6.12E-08 6.12E-08

Page 97: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

81

DFM Model of the Benchmark System

A DFM model of the BS represents the temporal behavior and dynamic evolution of the system.

To make an accurate representation of the system, vital components that dictates the system

behavior is identified and included in the system model. For instance, considering the case of the

BS, all the three (3) control units must be included in the model. Furthermore, each of the control

units are considered as a single component rather than a system. Once the vital parameters are

identified, causal, temporal and conditional relationships are established among the parameters.

These relationships are manifested by algebraic or differential equations, which can be solved to

construct the decision tables. For analysis purpose and model development, several assumptions

are made, including:

1. All assumptions made in the previous section applies;

2. Level sensor failure characteristics are omitted;

3. Unit fail-closed and fail-open are considered as sink states.

A stepwise representation of the methodology application and model development of the

benchmark system in the computer code DYMONDA is provided.

1. Identification of system hardware components

All hardware components are included in the DFM model and are represented by process variables

nodes in the model. The BS hardware components are presented in Table 16.

Table 16: System hardware identification

No. Description Representation

1. Unit-1 U1

2. Unit-2 (Standby unit) U2-SB

3. Unit-3 U3

4. Level Sensor ML

2. Identification of system parameters

Parameters that captures the attributes of the hardware components are identified and modelled as

process variables nodes in the DFM model. The BS parameters are presented in Table 17.

Page 98: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

82

Table 17: System parameter identification

No. Description Representation

1. Tank Level TL

2. Total Net Fluid Flow TNF

3. Total Liquid Inflow TIF

4. Total Liquid Outflow TOF

5. Liquid Flow from Unit-1 U1-F

6. Liquid Flow from Unit-2 U2-F

3. Identification of Condition Nodes

A condition node, like the process variable nodes represents physical parameters. However,

conditioning nodes more explicitly identify the switching action or failure modes/states of a

component. Condition nodes of the BS are presented in Table 18.

Table 18: System conditioning node identification

No. Description Representation

1. Unit-1 State or failure modes U1-S

2. Unit-2 State or failure modes U2-S

3. Unit-3 State or failure modes U3-S

4. Relationship Establishment (causal, temporal and conditional)

The process variable nodes are linked together by causality edges through transfer boxes and

transition boxes to model the cause-effect and temporal relationships among the parameters. The

conditioning nodes are linked to the transfer or transition box by the condition edges. For example,

consider Figure 19: the total liquid flow from Unit-1 (U1-F) is dependent on the command signal

to Unit-1 (U1) and the failure modes of Unit-1 (U1-S) which are stochastic in nature. U1 and U1-

F have a causal relationship which are represented and linked by causality edges, whereas U1-S

and U1-F have a conditional relationship which is represented and linked by a condition edge (as

can be seen from figure below).

Page 99: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

83

Figure 19: DFM model of Total Liquid Flow from Unit-1

The process of node linking via causality and condition edges is carried out for among all the

parameters having a causal, temporal and conditional relationships. Eventually, this result in an

integrated causality and time transition network as shown in Figure 20.

Figure 20: DFM Model of the Benchmark System

5. Discretization of system parameters

The identified process variable nodes and the condition nodes are then discretized into finite

number of states for the purpose of analysis. The number of discretized states or cells depends on

the analyst choice, level of in-depth modelling and the desired outcome. A sample discretization

of liquid level is into five (5) mutually exclusive intervals is depicted in Table 19.

Page 100: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

84

Table 19: Discretization of TL into 5 intervals

State Intervals/Cells Description

-2 Low-low (Sink cell) (𝑥 < −3)𝑚𝑒𝑡𝑒𝑟𝑠

-1 Low (𝑅𝑒𝑔𝑖𝑜𝑛:−1) (−3𝑡𝑜 − 1)𝑚𝑒𝑡𝑒𝑟𝑠

0 Nominal level (𝑅𝑒𝑔𝑖𝑜𝑛: 0) (−1𝑡𝑜 + 1)𝑚𝑒𝑡𝑒𝑟𝑠

+1 High (𝑅𝑒𝑔𝑖𝑜𝑛:+1) (+1𝑡𝑜 + 3)𝑚𝑒𝑡𝑒𝑟𝑠

+2 High-high (Sink cell) (𝑥 > +3)𝑚𝑒𝑡𝑒𝑟𝑠

6. Development of decision table

The decision tables within the transfer and transition boxes are constructed systematically via

inductive approach, i.e., enumerating all possible combinations of input states and then finding the

output state for each combinations. This ensures a complete and comprehensive table. Since

enumerating all possible input combinations can be time taking, decision table reduction method

is utilized to produce a compact table, suitable for use in the DYMONDA code. Sample decision

table is depicted in Table 20. Note: “*” represents a “Don’t care” condition.

Table 20: Decision table for TT1 with 5 discretized liquid level intervals

TNF @t=-1 TL @t=-1 TL @t=0

* -2 -2 (sink state)

* +2 +2 (sink state)

-1 -1 -2

-1 0 -1

-1 +1 0

0 -1 -1

0 0 0

0 +1 +1

+1 -1 0

+1 0 +1

+1 +1 +2

+2 -1 0

+2 0 +1

+2 +1 +2

Page 101: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

85

The decision table TT1 captures the time element in the model. Notice that there are two TL but

with different time element (@t= -1 and @t= 0). This implies that the current liquid level @t= 0

is dependent on the net liquid flow and the liquid level in the previous time step @t= -1. This

temporal transition relationship (system dynamics) is capture by the decision Table 20.

7. Model Analysis

Two scenarios were analyzed i.e., overflow and drained with two different discretization scheme:

(1) With 5 discretized state variable intervals; (2) With 8 discretized state variable intervals. The

top events in the model are defined in terms of state variable magnitudes, e.g., TL=+2 @t=0 and

TL=-2 @t=0. For instance, TL=+2 @t=0 is interpreted as the liquid level being above 3 meters at

time, t= 0. The simulation time step is chosen in such a way that transition occurs from an interval

to the adjacent interval in a single time step. This avoid multiple intervals transition and provide

smooth and finer results. Taking into account the number of discretized tank level, a single

simulation time step was approximated to be 1 hour. Once the top event and simulation time step

is defined, backtracking is performed through the model to generate timed PIs. For illustration

purpose, few generated results are presented below.

4.5.1. DFM Model results

A. For TL=+2 @t=0 or ‘overflow’ with single time step (5 intervals)

The first step is to define a top event in terms of the process variable node. The defined top event

is then expressed as a transition table. The depth of backtracking process is selected (analyst

choice). For the current case, a single time step is selected for demonstration purpose. A reduced

transition table is shown in Table 21.

Table 21: Transition table after eliminating U2-SB @t=-1

U3-S

@t= -1

U1-S

@t= -1

U2-S

@t= -1

TL

@t= -1

TOP TL=+2 @t=0

* * * +2 True

* +1 +1 +1 True

-1 * +1 +1 True

-1 +1 * +1 True

Page 102: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

86

The above transition table cannot be further reduced since the analysis is carried out only for a

single time step. Thus, the backtracking process and transition table expansion is terminated and

completed for the current analysis time step, t=-1. The transition table in Table 21 is the critical

transition table. A further absorption and reduction merging as well as consensus operation is

performed on the critical transition table to obtain the complete base, but they do not produce any

change in the critical transition table. Thus, Table 21 is the complete base for the top event. The

rows of Table 21 are the 4 PIs obtained for the Top Event “TL = +2 @ t = 0”. Quantification of

the above PIs can be carried out with ease by inputting failure rates of each units in the respective

process variable or conditioning nodes. The sum of the probabilities of mutually exclusive PIs

yields the exact probability of the top event. For illustration purpose, quantification of the top event

for system overflow with single time step is shown Table 22.

Table 22: Quantification of prime implicants for system overflow

# Prime Implicants Time Logic Probability

1. Tank Liquid Level was at +3 meters @t= -1 -

2. Tank Liquid Level was between +1 and

+3 meters

Unit-1 fail-open

Unit-3 fail-closed

@t= -1

@t= -1

@t= -1

AND

AND

6.52 × 10−6

3. Tank Liquid Level was between +1 and

+3 meters

Unit-2 fail-open

Unit-3 fail-closed

@t= -1

@t= -1

@t= -1

AND

AND

4.46 × 10−6

4. Tank Liquid Level was between +1 and

+3 meters

Unit-1 fail-open

Unit-2 fail-open

@t= -1

@t= -1

@t= -1

AND

AND

3.57 × 10−6

Top Event Probability 1.455 × 10−5

Page 103: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

87

B. For TL=+2 @t=0 with two time step

System analysis can be performed for the top event “TL=+2 @t=0” or “system overflow” as done

in the previous section, however with 2 simulation time steps or in other words for 2 hours in this

case. This implies that backtracking process will be carried out for 2 cycles. Complete base of the

top event “TL=+2 @t=0” with 2 simulation time steps is shown in Table 23.

Table 23: Complete base for the top event “TL=+2 @t=0” with 2 time steps

U1-S

@t= -1

U1-S

@t=-2

U2-S

@t= -1

U2-S

@t= -2

U3-S

@t= -1

U3-S

@t= -2

TL

@t= -1

TOP

* * * * * * +2 True

* * +1 0 -1 -1 +1 True

* * +1 +1 -1 -1 0 True

* * +1 +1 -1 -1 +1 True

* * +1 +1 -1 0 +1 True

* 0 +1 +1 -1 0 0 True

* 0 +1 0 -1 -1 0 True

+1 0 * * -1 -1 0 True

+1 0 * * -1 -1 +1 True

+1 +1 * * -1 -1 0 True

+1 +1 * * -1 -1 +1 True

+1 +1 * * -1 0 +1 True

+1 +1 +1 +1 * * 0 True

+1 0 +1 +1 * * +1 True

+1 +1 +1 0 * * +1 True

+1 +1 +1 +1 * * +1 True

+1 0 +1 +1 * * 0 True

The rows of Table 23 are the 17 prime implicants obtained for the top event “TL = +2 @ t = 0”

with 2 simulation time steps. The quantification of the PIs results in:

𝑇𝐿 = +2(@𝑡 = 0) = 7.698𝐸 − 05

Page 104: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

88

C. For ‘TL=+3 @t=0’ with 1 backtracking depth and 8 state variable intervals

The model analysis is performed with increased number of state variable intervals to check the

sensitivity of the PIs with respect to the number of states variable intervals. Table 24 presents the

PIs for system overflow with 8 state variable intervals for 1 backtracking depth.

Table 24: PIs for system overflow “TL=+3 @t= 0”

# Prime Implicants Time Logic

1. Tank Liquid Level was between +2 and +3 meters @t= -1 AND

Unit-1 Fail-High @t= -1 AND

Unit-2 Fail-High @t= -1

2. Tank Liquid Level was between +2 and +3 meters @t= -1 AND

Unit-1 Fail-High @t= -1 AND

Unit-3 Fail-Low @t= -1

3. Tank Liquid Level was between +2 and +3 meters @t= -1 AND

Unit-2 Fail-High @t= -1 AND

Unit-3 Fail-Low @t= -1

4. Tank Liquid Level was between +1 and +2 meters @t= -1 AND

Unit-1 Fail-High @t= -1 AND

Unit-2 Fail-High @t= -1 AND

Unit-3 Fail-Low @t= -1

System overflow [(𝑇𝐿 = +3(@𝑡 = 0)] 1.456𝐸 − 05

From Table 24, it may be observed that the PIs generated for the top event ‘system overflow’ with

state variable intervals of 5 (Case-I) differs from the one with 8 state variable intervals (Case-II).

Specifically, the prime implicant 4 (PI-4) is an additional PI. PI-4 does not appear in Case-I due to

the fact that it is eliminated during the simplification process by Boolean algebra or absorption

rule. Recall that PI-4 appears as a cut-set in the fault tree analysis, but not as a minimal cut-set.

Similarly, PI-4 appears as an implicant in Case-I, but not as a PI. This can be attributed to less

number of mutually exclusive state variable intervals for Case-I. Notice that in PI-4, the tank liquid

level was between +1 and +2 meters @t= -1”. This additional discretized interval of state variable

causes the PI-4 to be a timed-prime implicants in the complete base, even though a combination

Page 105: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

89

of “Unit-1 fail-high AND Unit-2 fail-high AND Unit-3 fail-low @t=-1” is not a PI. It is important

to note that, due to the dynamic evolution of the state variable with time and the dependencies of

the state variable on the unit state combination, a system overflow scenario is observed in a single

time step with PI-4. This is true since the volume of the discretized intervals covering the entire

state space for Case-I is larger than for Case-II. Hence, many transition from an interval will tend

to remain in the same interval given that the simulation time is small, and the volume of the

discretized intervals are large. However, for Case-II, the state variable can make a transition to

another discretized intervals due to the smaller interval volume.

Similarly, simulation for 2 backtracking depth generates 28 PIs for Case-II, whereas for Case-I

only 17 PIs are generated. The increased number of state intervals provide the analyst a more detail

information and also increases the accuracy of the result obtained. However, computation time and

complexity increase significantly with increased number of mutually exclusive intervals.

D. Variation of PIs for 5 and 8 state variable intervals

It may be observed that the number of PIs increases as the backtracking depth is increased. For

instance, the number of PIs increased from 4 to 28 as the depth increased from 1 to 2. This implies

that, with increasing depth there will be an increased number of possible ways for the system to

fail due to the availability of time for the units to make transitions from normal to failure state.

This analysis is performed with 5 and 8 state variable intervals to check the sensitivity of the

number of PIs with respect to variation in number of state variable intervals. The sensitivity plot

is depicted in Figure 21 for 10 backtracking depth. Of course, the state variable can be discretized

into a larger number of intervals to obtain a finer result. However, system modelling and

development of decision table becomes complex with increased number of discretized intervals.

Page 106: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

90

Figure 21: Sensitivity of number of PIs to discretized state variable intervals

4.5.2. Timed-Fault Tree Generation from DFM

For illustration and discussion purpose, the generation of timed-fault trees for system overflow

scenario with 5 discretized state variable intervals is presented below. Similar to the classical FT

technique, the first step is to define a desired top event in terms of state variable magnitude i.e.,

“TL=+2 @t=0” which is associated with the transition table “TT1”. The decision table of TT1 is

utilized to determine the inputs that causes the tank liquid level to be in TL=+2 @t=0. Since the

TT1 associated with time delays, it changes the time at which a particular variable state occurs.

The first transition from time 𝑡 = 0 to 𝑡 = −1 can be observed in the timed-fault tree given below

(See Figure 22). Throughout the backtracking process, dynamic and physical consistency rules are

applied to process variables and conditioning nodes similar to that applied while developing the

DFM model.

Page 107: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

91

Figure 22: Timed-fault tree top event and its transition from @t=0 to @t= -1

It can be observed in Figure A-III-1 (Appendix III) that the backtracking process from the top event

“TL= +2 @t=0” results in the tank level TL= -1 @t= -1 or TL= -2 @t= -1 (G28 and G33).

However, the tank level TL= +1 @t= -1 have already occurred under the same parent AND gate

(G3 in Figure 22). This condition is impossible due to physical consistency i.e., tank liquid level

cannot assume two different values/state in the same time step. Hence, G28 and G33 is eliminated

from the tree. This elimination further leads to pruning of the parent AND gate G25.

In Figure A-III-2 (transfer-in gate G7), the backtracking process for single time step @𝑡 = −1

leads to tank level being at 𝑇𝐿 = 0 ∪ −1 ∪ −2@𝑡 = −1 (G37). However, under the same parent

Page 108: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

92

AND gate (G2), 𝑇𝐿 = +2@𝑡 = −1 has already occurred. By physical consistency check, this

condition is impossible to occur since liquid level cannot be in two or more different state in the

same time step. This results in an elimination of gate G37. The removal of G37 further lead to a

pruning of the AND gate G35, since initial pruning of G37 made the succeeding event occurrence

to be impossible. The same argument holds for G46 and G48 in the transfer gate G8 (See Figure

A-III-3).

Application of physical and dynamic consistency rule reduces the overall timed-fault tree to Figure

A-III-6: . The FT in Figure A-III-4 is further pruned by application of physical consistency rule to

generate the final timed-FT shown in Figure 23. Since the Unit-1 fail-high have already occurred

in G7, Unit-1 fail-closed (G23) cannot occur due to physical consistency, i.e., the Unit-1 cannot

be in two different state in the same time step. Similarly, this argument holds for Unit-2 state. Unit-

2 fail-high has already occurred in G8, hence Unit-2 fail-closed (G53) cannot occur due to physical

consistency, i.e., Unit-2 cannot be in two different state in the same time step. These pruning are

due to the fact that the gates G7 and G8 first occurs under the AND gate G6 rather than the AND

gate G14. The pruning would have been reversed if gate G14 appears first in the timed-FT then

gate G6, i.e., G7 and G8 would have been pruned, whereas G23 and G53 would have appeared in

the MCS. The MCSs for the final reduced timed-fault tree are given in Table 25.

Table 25: Minimal cut-set for the final timed-fault tree

1. TL= +2 @t= –1 OR

2. (TL= +1 AND U1-S= +1 AND U2-S= +1) @t= –1 OR

3. (TL= +1 AND U1-S= +1 AND U3-S= –1) @t= –1 OR

4. (TL= +1 AND U2-S= +1 AND U3-S= –1) @t= –1

It can be observed that the MCS for the timed-fault tree is exactly the same as the PIs generated

using the DFM model. In other words, this process of backtracking and generation of time-fault

trees from the DFM model points out the fact that results obtained from dynamic methods can be

incorporated into an existing static PRA model.

Page 109: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

93

Figure 23: Final timed-fault tree after consistency check

Page 110: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

94

Markov-CCMT Model of the Benchmark System

The classical Markov model of the benchmark system presented provide a time-dependent

behavior of the system via a transition matrix, however lacks capability to capture system

dynamics. This section couples the classical Markov model with CCMT.

a. Computation of ℎ(𝑛 𝑛′⁄ , 𝑘∆𝑡)

The state transition matrix can be developed considering assumption number 4 in Sub-section

3.2.3, in that the units have a statistically independent failure. The unit stochastic behavior is

represented through ℎ(𝑛 𝑛′⁄ , 𝑘∆𝑡), which is the probability that the unit state combination at time

𝑡 = (𝑘 + 1)∆𝑡 is at 𝑛, given that the unit state at time 𝑡 = 𝑘∆𝑡 is at 𝑛′. It can be expressed as:

ℎ(𝑛 𝑛′⁄ , (𝑘 + 1)∆𝑡) = ℎ{𝑛((𝑘 + 1)∆𝑡) = 𝑛 𝑛(𝑘∆𝑡) = 𝑛′}⁄

(4-11)

It can be observed that ℎ(𝑛 𝑛′⁄ , 𝑘∆𝑡) is a conditional probability of the component state

combination being at 𝑛 at time 𝑡 = (𝑘 + 1)∆𝑡, given that the component state combination was at

𝑛′ at time 𝑡 = 𝑘∆𝑡. ℎ(𝑛 𝑛′⁄ , 𝑘∆𝑡) can be expressed as in Equation (3-50).

For instance, 𝑛1 = 0, 𝑛2 = 0𝑎𝑛𝑑𝑛3 = 0 at time 𝑘∆𝑡 implies that all the three units are normal,

and hence 𝑛′ = 0. If 𝑛1 = 0, 𝑛2 = 0𝑎𝑛𝑑𝑛3 = 0 at time (𝑘 + 1)∆𝑡 i.e., Unit-1, Unit-2 and Unit-

3 are normal, which implies 𝑛 = 0 (as per Table 12). Thus ℎ(𝑛 𝑛′⁄ , (𝑘 + 1)∆𝑡) can be written as;

ℎ(0 0⁄ , (𝑘 + 1)∆𝑡) = (𝑛1 = 0 𝑛′1 = 0⁄ ) ∗ (𝑛2 = 0 𝑛′2 = 0⁄ ) ∗ (𝑛3 = 0 𝑛′3 = 0⁄ ) (4-12)

i.e., a transition of the system from normal state to normal state at time step 𝑘∆𝑡, [(𝑘 + 1)∆𝑡].

ℎ(0 0⁄ , (𝑘 + 1)∆𝑡) can be quantified in terms of the unit failure probabilities:

ℎ(0 0⁄ , (𝑘 + 1)∆𝑡) = (1 − 𝜆1∆𝑡) ∗ (1 − 𝜆2∆𝑡) ∗ (1 − 𝜆3∆𝑡) (4-13)

ℎ(0 0⁄ , (𝑘 + 1)∆𝑡) = 0.987

The transition probabilities among the other state combinations can be computed in the same

fashion, and the general form is given in Table 26.

Page 111: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

95

Table 26: System state transition probabilities

ℎ(𝑛 𝑛′⁄ , (𝑘 + 1)∆𝑡) = 𝑐𝑚(𝑛𝑚 𝑛′𝑚⁄ , 𝑘∆𝑡)

𝑀

𝑚=1

Unit-1 (𝑛1 𝑛′1⁄ , 𝑘∆𝑡) Unit-2 (𝑛2 𝑛′2⁄ , 𝑘∆𝑡) Unit-3 (𝑛3 𝑛′3⁄ , 𝑘∆𝑡)

𝑃(𝑛1 = 0 𝑛′1⁄ = 0)

= (1 − (𝜆1𝑓𝑐+ 𝜆1

𝑓𝑜)∆𝑡)

𝑃(𝑛2 = 0 𝑛′2⁄ = 0)

= (1 − (𝜆2𝑓𝑐+ 𝜆2

𝑓𝑜)∆𝑡)

𝑃(𝑛3 = 0 𝑛′3⁄ = 0)

= (1 − (𝜆3𝑓𝑐+ 𝜆3

𝑓𝑜)∆𝑡)

𝑃(𝑛1 = 1 𝑛′1⁄ = 0) = 𝜆1𝑓𝑐∆𝑡 𝑃(𝑛2 = 1 𝑛′2⁄ = 0) = 𝜆2

𝑓𝑐∆𝑡 𝑃(𝑛3 = 1 𝑛′3⁄ = 0) = 𝜆3

𝑓𝑐∆𝑡

𝑃(𝑛1 = 2 𝑛′1⁄ = 0) = 𝜆1𝑓𝑜∆𝑡 𝑃(𝑛2 = 2 𝑛′2⁄ = 0) = 𝜆2

𝑓𝑜∆𝑡 𝑃(𝑛3 = 2 𝑛′3⁄ = 0) = 𝜆3

𝑓𝑜∆𝑡

𝑃(𝑛1 𝑛′1⁄ ) = 0, otherwise 𝑃(𝑛2 𝑛′2⁄ ) = 0, otherwise 𝑃(𝑛3 𝑛′3⁄ ) = 0, otherwise

Since there are 27 unit state combinations, ℎ(𝑛 𝑛′⁄ , (𝑘 + 1)∆𝑡) will have a 27 × 27 dimension

matrix with the structure like Table 27. The algorithm for generating the overall matrix is

developed and implemented using Fortran95. Note that, ℎ(𝑛 𝑛′⁄ , (𝑘 + 1)∆𝑡) is expressed as a

product of individual unit failure probabilities due to the initial assumption made for the BS that

unit failures are statistically independent.

Table 27: Possible state transition among distinct unit state combinations

𝒏′ 𝒏

1 2 3 ……….. 𝒊 ……….. 𝒏

1 𝜇11∆𝑡 𝜆12∆𝑡 𝜆13∆𝑡 ……….. 𝜆1𝑖∆𝑡 ……….. 𝜆1𝑛∆𝑡

2 0 𝜇22∆𝑡 𝜆23∆𝑡 ……….. 𝜆2𝑖∆𝑡 ……….. 𝜆2𝑛∆𝑡

⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮

𝒋 0 0 0 ……….. 𝜇𝑗𝑖∆𝑡 ……….. 𝜆𝑗𝑛∆𝑡

⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮

𝒏′ 0 0 0 0 0 ……….. 𝜇𝑛′𝑛∆𝑡

For the BS under consideration, the conditional transition probability matrix ℎ(𝑛 𝑛′⁄ , 𝑘∆𝑡)

represent the BS stochastic time-dependent properties. A further analysis of the transition

Page 112: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

96

probability matrix can be performed by decomposing the matrix into four (4) sub-matrices and re-

writing the transition probability matrix ‘P’ in the canonical form as shown in Table 28.

Table 28: Canonical form of the transition probability matric

P 𝑺′ 𝑻′

𝑺 𝑃𝑒𝑟𝑠𝑖𝑠𝑡𝑒𝑛𝑡𝑠𝑡𝑎𝑡𝑒𝑠(𝐼) 𝑇𝑟𝑎𝑛𝑠𝑖𝑒𝑛𝑡𝑠𝑡𝑎𝑡𝑒𝑠(𝑇)

𝑻 𝑍𝑒𝑟𝑜𝑠𝑡𝑎𝑡𝑒𝑠(0) 𝑇𝑟𝑎𝑛𝑠𝑖𝑒𝑛𝑡𝑠𝑡𝑎𝑡𝑒𝑠(𝑄)

The decomposition is performed to analyze the evolution of the system from a transient state by

making use of the transient sub-matrices 𝑇 and 𝑄. This enable one to understand the steady state

of the system, or the probability that the system will eventually end up in a specific sink system

state. Hence, for analysis purpose, layer-3 of the Markov model i.e., all unit failure is considered

as the sink state of the system. This consideration is valid since as time 𝑡 → ∞, the system will end

up in any of the states within layer-3. The sub-matrices 𝐼 and 𝑄 are square matrix, and 𝑇 are

rectangular matrix. Hence, if the total number of states in the system is 𝑛 states and the number

of absorbing states is 𝑚, then the total number of non-absorbing states is (𝑛 − 𝑚) states. It is

known that the sub-matrix 𝐼 is composed of absorbing states since it is an identity matrix, and thus

have the dimension of (𝑚 ×𝑚) matrix. Since 𝑄 is a non-absorbing transient square sub-matrix,

its dimension is [(𝑛 − 𝑚) × (𝑛 −𝑚)] matrix, which lead us to the matrices 𝑇 and 0 having a

dimension of [𝑚 × (𝑛 −𝑚)] and [(𝑛 − 𝑚) ×𝑚] respectively. It is obvious from the above matrix

decomposition that the transient group (𝑇) describe the transition of states from transient group to

persistent groups. Whereas, the transient group (𝑄) describe the transition of system states from

transient group (𝑄) to transient group (𝑇). The transient sub-matrices are extracted from the main

transition probability matrix for further analysis of matrices T and Q. The fundamental matrix (N)

of P can be determined using:

𝑁 = (𝐼 − 𝑄)−1

(4-14)

Key information can be extracted from the matrix N. The sum of the elements of the 𝑗𝑡ℎ column

of N gives the expected absorption time of the 𝑗𝑡ℎ column transient state to be absorbed into a

persistent states (See Table 29). For example, taking the last column of the fundamental matrix 𝑁,

Page 113: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

97

the sum of the elements is ≅ 448, which is the expected (mean) absorption time of the transient

state 0 (all units normal) into any of the persistent states conditioned upon that the system started

at state 0 (initial condition). A further analysis can be performed by considering the individual

elements of the last column vector in matrix N. For instance, considering the last element of the

column vector, which is ≅ 75, it can be interpreted as the mean number of time the system is in

state-0 given that the system started in state-0 initially. Or, considering the first element which is

≅ 17; it can be interpreted as the number of time the system is in state-18 given that the system

was initially in State-0. Similarly, other elements of the transition matrix can be explained.

The matrix TN gives the steady-state probabilities for ending in any absorbing state given that the

system started in a transient state. The (𝑖, 𝑗)𝑡ℎelements of the transient matrix 𝑇𝑁 is the probability

of being absorbed into persistent or absorbing state 𝑖 from the transient state 𝑗. This absorption

probability of the transient states is represented by 𝛼𝑗𝑖. For illustration purpose considering the

matrix element 𝛼0,26 = 0.125, it can be interpreted as the probability of transition from the

transient State-0 to the absorbing State-26. Or, considering the column vector of TN 𝛼0,𝑖 = 0.125,

which means if the system started at State-0 (all units normal), it is equally likely for the system

to end up in any of the absorbing states. Again, consider the first column of the matrix TN. It can

be observed that 𝛼18,26 = 𝛼18,22 = 0.5, i.e., the probability of the system being absorbed in state-

26 given that the system was in state-18 is equally likely, whereas the other elements 𝛼18,𝑗 is zero.

This can be validated by going back to the Markov model state representation of the system, i.e.,

if the system is in state-18 (𝑛1 = 0, 𝑛2 = 2, 𝑛3 = 2), it can only make a transition to either State-

22 or 26 (𝑛1 = 1𝑜𝑟2, 𝑛2 = 2, 𝑛3 = 2). This is due to the fact that the BS is a non-repairable

system. The matrix TN can be validated as the columns of the matrix must always be equal to 1.00.

Page 114: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

98

Table 29: Fundamental matrix 𝑁 = (𝐼 − 𝑄)−1

Page 115: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

99

Table 30: Matrix TN

TN matrix

Absorbing

System

States (𝑖)

Transient System States (𝑗)

18 17 16 15 14 13 12 11 10 9 8 7 6 5 4 3 2 1 0

26 0.5 0 0 0 0.5 0 0.5 0 0 0 0 0 0.25 0 0.25 0 0.25 0 0.125

25 0 0.5 0 0 0 0.5 0.5 0 0 0 0 0 0 0.25 0.25 0 0.25 0 0.125

24 0 0 0.5 0 0.5 0 0 0.5 0 0 0 0 0.25 0 0 0.25 0.25 0 0.125

23 0 0 0 0.5 0 0.5 0 0.5 0 0 0 0 0 0.25 0 0.25 0.25 0 0.125

22 0.5 0 0 0 0 0 0 0 0.5 0 0.5 0 0.25 0 0.25 0 0 0.25 0.125

21 0 0.5 0 0 0 0 0 0 0 0.5 0.5 0 0 0.25 0.25 0 0 0.25 0.125

20 0 0 0.5 0 0 0 0 0 0.5 0 0 0.5 0.25 0 0 0.25 0 0.25 0.125

19 0 0 0 0.5 0 0 0 0 0 0.5 0 0.5 0 0.25 0 0.25 0 0.25 0.125

Sum 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1

Page 116: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

100

Limitations of the time-dependent Markov chain can be observed after a detail analysis of state

conditional probability matrix, and also recalling the system qualitative analysis performed in

Section 4.4. Note that, these limitations were underlined with the knowledge of dynamic PSA.

• For the classical Markov chain, it was assumed that the system states 19 to 26 are absorbing

states, however this is not true from the perspective of dynamic PRA, i.e., states 19, 22 and

24 that belongs to the set of absorbing states are not necessarily absorbing since these states

do not lead to system overflow or drained scenario.

• The system states 12, 13 and 17 are assumed as transient states in the classical Markov

model, however this in not true from a dynamic PRA standpoint since these states lead to

system overflow and drained condition. Thus, the system states 12, 13 and 17 must be

considered as an absorbing states for dynamic PRA.

• The most important aspect to be taken from this analysis is that, the classical Markov model

is concerned with only the system hardware states thus neglecting the dynamic evolution

of the system. System failure is explicitly defined only on system configuration, i.e., system

failure occurs when all the 3 units fail. This is not necessarily true since the system can end

up in a quasi-stable state even if all the 3 units’ fail (the phenomenon was observed in event

tree analysis).

The conditional transition probabilities matrix ℎ(𝑛 𝑛′⁄ , 𝑘∆𝑡) generated can be coupled with

system dynamics to generate a joint system probability matrix, and hence providing a possible

approach to perform an integrated deterministic and probabilistic system assessment.

b. Computation of 𝑔(𝑗 𝑗′,⁄ 𝑛′, 𝑘∆𝑡)

For determination of 𝑔(𝑗 𝑗′,⁄ 𝑛′, 𝑘∆𝑡), the sate space is first discretized into finite number of

computational cells. The equal weight quadrature scheme was used to partition the state variable

magnitude. It is important to note that the transition among discretized cells depends on both the

initial cell location and the system configuration at that instant. Thus, care must be taken in

defining the initial location of the state variable within a cell. For the BS, it is obvious that the

controlled variable state space is one dimensional i.e., liquid level (𝑥). CCMT requires the

knowledge of top events in terms of state variable magnitude or computational absorbing cells.

Discretization scheme of the state variable along with the control regions is shown in Table 31.

Page 117: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

101

Table 31: Controlled variable discretization scheme

𝑉𝑟 Variable

state (𝐽) Variable magnitude

range 𝑉𝑗

System

State

- 𝑗 = 1 𝑥 < −3𝑚 𝑉1 Drained

𝑉𝑟 = 1 𝑗 = 2 −3𝑚 ≤ 𝑥 ≤ −1𝑚 𝑉2 Low

𝑉𝑟 = 2 𝑗 = 3 −1𝑚 ≤ 𝑥 ≤ +1𝑚 𝑉3 Normal

𝑉𝑟 = 3 𝑗 = 4 +1𝑚 ≤ 𝑥 ≤ +3𝑚 𝑉4 High

- 𝑗 = 5 𝑥 > +3𝑚 𝑉5 Overflow

Note: The top events are shaded (green) in the table, i.e., overflow and drained

The three (3) control regions 𝑉𝑟 can be further divided into arbitrary number of sub-cells depending

on the analyst choice and the outcome desired. For the benchmark system, since Vr are disjoint

intervals themselves, 𝑉𝑗 are identical to Vr for minimum 𝐽𝑟 (i.e., 𝐽𝑟 = 1) and partitioning is unique

with a single departure point 𝑃 = 1 from each region. Of course, more number of sub-cells

increases the result accuracy. However, computational complexity and intensity increases with

increased number of sub-cells/cells. A possible equal weight quadrature scheme with increased

number of sub-cells is shown in Table 32. For the current analysis, 𝑃 = 3 is chosen for

demonstration of the methodology at the same time balancing the accuracy and computation

complexity. It can be observed from Table 32 that there are three departure points 𝑃 = 3 from

each sub-cells, or the initial liquid level in the tank which is an input for system simulation.

Table 32: Cell discretization via equal weight quadrature scheme

𝐽 Liquid Level (𝑥) Equal weight Quadrature scheme

𝑃 = 1 𝑃 = 3 𝑃 = 5

𝑗 = 1 𝑥 < −3𝑚 - - -

𝑗 = 2 −3𝑚 ≤ 𝑥 ≤ −1𝑚 −2𝑚

−2.5𝑚 −2.75𝑚

−2.25𝑚

−2.0𝑚 −2.0𝑚

−1.5𝑚 −1.75𝑚

−1.25𝑚

𝑗 = 3 −1𝑚 ≤ 𝑥 ≤ +1𝑚 0𝑚 −0.5𝑚 −0.75𝑚

Page 118: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

102

−0.25𝑚

0𝑚 0𝑚

+0.5𝑚 +0.25𝑚

+0.75𝑚

𝑗 = 4 +1𝑚 ≤ 𝑥 ≤ +3𝑚 +2𝑚

+1.5𝑚 +1.25𝑚

+1.75𝑚

+2.0𝑚 +2.0𝑚

+2.5𝑚 +2.25𝑚

+2.75𝑚

𝑗 = 5 𝑥 > +3𝑚 - - -

The state space discretization can be fairly mechanized by implementing it in a computer code

given that the control space and number of sub-cells are defined. The source code for state space

discretization is shown in the Appendix IV. Once the control space is discretized into sub-spaces,

the trajectories of the state vector in the discretized space must be determine. For the benchmark

system, the dynamic evolution of the liquid level in the tank is defined by:

𝑑𝑥(𝑡)

𝑑𝑡= 𝑓𝑛

(4-15)

Or, Equation (4-15) can be re-write as:

𝑑𝑥(𝑡) = 𝑓𝑛 ∗ 𝑑𝑡

𝑥(𝑡 + ∆𝑡) − 𝑥(𝑡) = 𝑓𝑛 ∗ ∆𝑡 (4-16)

Thus, liquid level in the next time step is given by:

𝑥(𝑡 + ∆𝑡) = 𝑥(𝑡) + 𝑓𝑛 ∗ ∆𝑡 (4-17)

Here, we can re-define the above equation for our analysis. It is considered here that:

𝑥(𝑡 + ∆𝑡) = 𝑥(𝑡) + 𝑓𝑛 ′ ∗ ∆𝑡

(4-18)

The above system dynamics is written in 𝑛′ terms due to the assumption in the methodology that

the component state combination does not change for a small-time step ∆𝑡, i.e., the computation

Page 119: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

103

of 𝑔(𝑗 𝑗′,⁄ 𝑛′, 𝑘∆𝑡) is conditioned upon that 𝑛′ does not change in the small time step ∆𝑡. In fact,

𝑔(𝑗 𝑗′,⁄ 𝑛′, 𝑘∆𝑡) itself constitute a Markov chain. Since all the parameters on the right hand side of

the Equation 42 is known or can be defined by the user, 𝑥(𝑡 + ∆𝑡) can be determined with ease

for any given time step. The dimension of the state vector trajectories 𝑥(𝑡 + ∆𝑡) is dependent on

the number of departure points 𝑥(𝑡), with both having the same dimension, and hence 𝑓𝑛′ ∗ ∆𝑡.

Once 𝑥(𝑡 + ∆𝑡) is computed from the system code, 𝑔(𝑗 𝑗′,⁄ 𝑛′, 𝑘∆𝑡) can be determined by

integrating over the control space.

For illustration purpose, first consider the control region-2 which is partitioned into 3 equally

distributed sub-cells. Hence, there are three trajectory segments departing from 𝑉2 at time 𝑡 with

the same 𝑛′, and arriving in IC number of 𝑉𝑗 at (𝑡 + ∆𝑡), depending on location of the controlled

variables within 𝑉2 which is the 𝑉𝑗′ for this case. For the benchmark system at state 𝑛′(𝑡) = 1, if

the trajectory segments depart from 𝑉21, 𝑉22𝑎𝑛𝑑𝑉23 at time 𝑡, then the arrival points 𝑉𝑗 at (𝑡 +

∆𝑡) and hence 𝑔(𝑗 𝑗′,⁄ 𝑛′, 𝑘∆𝑡) can be determined using the below algorithm. Since initially the

state variable is in 𝑉21, 𝑉22𝑎𝑛𝑑𝑉23 sub-cell which are within the cell 𝑉2 at time 𝑡, the total

probability of 𝑃𝑟𝑗,𝑛′(𝑡) must be:

𝑃21,1(𝑡) + 𝑃22,1(𝑡) + 𝑃23,1(𝑡) = 1 (4-19)

Here;

𝑃𝑟𝑗,𝑛′(𝑡) = 𝑃21,1(𝑡) = 𝑃𝑟{𝑛(𝑡) = 1, 𝑥(𝑡) ∈ 𝑉21 𝑥(𝑡) ∈ 𝑉2⁄ } (4-20)

𝑃22,1(𝑡) = 𝑃𝑟{𝑛(𝑡) = 1, 𝑥(𝑡) ∈ 𝑉22 𝑥(𝑡) ∈ 𝑉2⁄ } (4-21)

𝑃23,1(𝑡) = 𝑃𝑟{𝑛(𝑡) = 1, 𝑥(𝑡) ∈ 𝑉23 𝑥(𝑡) ∈ 𝑉2⁄ } (4-22)

By definition or by equal weight quadrature discretization scheme, the sub-cells 𝑉21, 𝑉22𝑎𝑛𝑑𝑉23

have equal volumes, i.e.,

𝑃21,1(𝑡) = 𝑃22,1(𝑡) = 𝑃23,1(𝑡) =1

3 (4-23)

Hence, the transition from 𝑉1 to 𝑉1 can be determined as follow:

𝑔(𝑉2 𝑉2⁄ , 1) = 𝑔(𝑉2 𝑉21⁄ , 1). 𝑃21,1(𝑡) + 𝑔(𝑉2 𝑉22⁄ , 1). 𝑃22,1(𝑡)

+𝑔(𝑉2 𝑉23⁄ , 1). 𝑃23,1(𝑡) (4-24)

𝑔(𝑉2 𝑉2⁄ , 1) = (1 ∗ 0.333) + (1 ∗ 0.333) + (1 ∗ 0.333)

Page 120: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

104

𝑔(𝑉2 𝑉2⁄ , 1) = 1.00

And hence 𝑔(𝑉𝑗 𝑉2⁄ , 1) = 0 for 𝑗 ≠ 1, since the state vector trajectories will remain in region-2

will 100% probability rather than making a transition to the other control regions. For the second

case, consider for 𝑛′ = 12 with the state vector depart from 𝑉21, 𝑉22𝑎𝑛𝑑𝑉23 at time 𝑡 which is

represented by 𝑉𝑗′ . The state vector trajectories and the arrival point 𝑉𝑗 can be computed a follow.

By equal weight quadrature discretization scheme:

𝑃21,12(𝑡) = 𝑃22,12(𝑡) = 𝑃23,12(𝑡) =1

3 (4-25)

The transition from 𝑉2 to 𝑉2 for 𝑛′ = 12 can be determined as:

𝑔(𝑉2 𝑉2⁄ , 12) = 𝑔(𝑉2 𝑉21⁄ , 12). 𝑃21,12(𝑡) + 𝑔(𝑉2 𝑉22⁄ , 12). 𝑃22,12(𝑡)

+𝑔(𝑉2 𝑉23⁄ , 12). 𝑃23,12(𝑡) (4-26)

𝑔(𝑉1 𝑉1⁄ , 12) = (1 ∗ 0.333) + 0 + 0

𝑔(𝑉2 𝑉2⁄ , 12) = 0.333

For transition from 𝑉2 to 𝑉3, we have:

𝑔(𝑉3 𝑉2⁄ , 12) = 𝑔(𝑉3 𝑉21⁄ , 12). 𝑃21,12(𝑡) + 𝑔(𝑉3 𝑉22⁄ , 12). 𝑃22,12(𝑡)

+𝑔(𝑉3 𝑉13⁄ , 12). 𝑃23,12(𝑡) (4-27)

𝑔(𝑉3 𝑉2⁄ , 12) = 0 + (1 ∗ 0.333) + (1 ∗ 0.333)

𝑔(𝑉3 𝑉2⁄ , 12) = 0.667

Hence, 𝑔(𝑉𝑗 𝑉2⁄ , 1) = 0 for 𝑗 ≠ 2, 3; since the state vector trajectories will remain in region-2

with probability of 33 % and in region-3 with a probability of 66.7 %. Similarly, the rest of the

trajectories can be computed mechanically via computer codes as shown in the Table 33.

𝑃𝑟{𝑔(𝑗 𝑗′, 𝑛′, 𝑘∆𝑡⁄ )} is a 5 × 5 matrix for each unit state combinations, which results in 135 × 5

dimensional matrix for 27 unit state combination.

Page 121: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

105

Table 33: 𝑃𝑟{𝑔(𝑗 𝑗′, 𝑛′, 𝑘∆𝑡⁄ )} for 𝑘 = 1

SI.No. 𝑛′

From/to 𝑗 Sum

𝒋′ 1 2 3 4 5

1 0 1 1 0 0 0 0 1.00

2 0 2 0 1 0 0 0 1.00

3 0 3 0 0 1 0 0 1.00

4 0 4 0 0 0 1 0 1.00

5 0 5 0 0 0 0 1 1.00

⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮

⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮

131 26 1 1 0 0 0 0 1.00

132 26 2 0 0.33 0.67 0 0 1.00

133 26 3 0 0 0.33 0.67 0 1.00

134 26 4 0 0 00 0.33 0.67 1.00

135 26 5 0 0 0 0 1 1.00

c. Computation of transition probability matrix 𝑞𝑛,𝑗𝑛′,𝑗′(𝑘∆𝑡)

The elements of the overall system transition matrix 𝑞𝑛,𝑗𝑛′,𝑗′(𝑘∆𝑡) is a function of both the cell to

cell transition probabilities 𝑔(𝑗 𝑗′, 𝑛′, 𝑘∆𝑡⁄ ) and the conditional unit state transition probabilities

ℎ(𝑛 𝑛′⁄ , 𝑗′ → 𝑗, 𝑘∆𝑡). 𝑞𝑛,𝑗𝑛′,𝑗′(𝑘∆𝑡) can be computed as follow:

𝑞𝑛 ,𝑗𝑛 ′ ,𝑗 ′ (𝑘∆𝑡) = 𝑔(𝑗 𝑗′ ,𝑛′ ,𝑘∆𝑡⁄ ).ℎ(𝑛 𝑛′⁄ , 𝑗′ → 𝑗,𝑘∆𝑡)

(4-28)

The sum of the column elements of 𝑞𝑛,𝑗𝑛′,𝑗′

(∆𝑡) must be equal to 1.00. Hence, the transition matrix

can be verified and validated once obtained before performing further computation. Since there

are 27 distinct system states and 5 cells of the controlled variable state space, the 𝑞𝑛,𝑗𝑛′,𝑗′(𝑘∆𝑡) matrix

will have an overall matrix dimension of 27 × 5 = 135 rows and 135 columns, where 𝑁 = 27

and 𝐽 = 5. Thus, the transition matrix 𝑞𝑛,𝑗𝑛′,𝑗′(𝑘∆𝑡) = 135 × 135 square matrix (a small portion of

which shown in Table 34).

Page 122: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

106

Table 34: A small portion of the overall system transition matrix 𝑞(𝑛, 𝑗 𝑗′⁄ , 𝑛′, 𝑘∆𝑡) for 𝑘 = 1

Page 123: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-107

d. Computation of 𝑃𝑛,𝑗((𝑘 + 1)∆𝑡)

The computation of 𝑃𝑛,𝑗(𝑘∆𝑡) for any specific time step can be performed using the inductive as

well as the deductive algorithm. For the current case, an inductive algorithm has been implemented

with the initial condition 𝑃𝑛′,𝑗′(0) to determine 𝑃𝑛,𝑗((𝑘 + 1)∆𝑡). For analysis purpose, the system

is assumed to be in normal condition with all the units being in normal state and the liquid level in

the nominal region. This can be quantitively expressed as;

𝑃𝑛 ′ ,𝑗 ′ (0) = 1 ; 𝑓𝑜𝑟 𝑛′ = 1, 𝑗′ = 30 ; 𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒

(4-29)

Or, the initial system condition is expressed as:

𝑃1,1(0) = 0;𝑃1,2(0) = 0;𝑃1,3(0) = 1;𝑃1,4(0) = 0;𝑃1,5(0) = 0;𝑃2,1(0) = 0;… . . 𝑃𝑁,𝐽(0) = 0

Once the initial condition and the time step are fixed, 𝑃𝑛,𝑗((𝑘 + 1)∆𝑡) can be computed utilizing

𝑞(𝑛, 𝑗 𝑗′⁄ , 𝑛′, 𝑘∆𝑡) and by implementing the Forward algorithm. In order to determine the top event

probabilities at any given time step, we must recall the space discretization scheme in which 𝑗 = 1

and 𝑗 = 5 are considered as absorbing cells in the system. Hence, the top events are described in

terms of the value of computational cells 𝑗. More specifically, 𝑗 = 1 implies liquid level 𝑥 < −3𝑚

which is the Drained condition, and 𝑗 = 5 implies 𝑥 > +3𝑚 which is the overflow condition.

Thus, for a specific time step, summing up all the elements of 𝑃𝑛,𝑗((𝑘 + 1). ∆𝑡) for 𝑗 = 1 and 𝑗 =

5 will give the probabilities of system drained and overflow scenario respectively. If the overall

probability of system failure (overflow + drained) at any given time step is represented by

𝑃𝑂𝐷(𝑘∆𝑡), it can be written as the summation of:

𝑃𝑂𝐷(𝑘∆𝑡) = 𝑃𝑛 ,1(𝑘∆𝑡)

𝑁

𝑛=1

+ 𝑃𝑛 ,5(𝑘∆𝑡)

𝑁

𝑛=1

(4-30)

Hence the system reliability 𝑅(𝑘∆𝑡) is given by (See Table 35):

𝑅(𝑘∆𝑡) = 1 − 𝑃𝑂𝐷(𝑘∆𝑡) (4-31)

Page 124: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-108

Table 35: Top event probabilities for the benchmark system

Time step 𝑃𝑛,1(𝑘∆𝑡) 𝑃𝑛,5(𝑘∆𝑡) 𝑃𝑂𝐷(𝑘∆𝑡)

𝑘 = 0 0 0 0

𝑘 = 1 0 0 0

𝑘 = 2 0 1.02E − 08 1.02E − 08

𝑘 = 3 1.86E − 07 9.17E − 05 9.19E − 05

𝑘 = 4 2.15E − 06 5.11𝐸 − 04 5.14E − 04

Table 36: Sample state transition probabilities

𝑷𝒏,𝒋 𝑷(𝒌 = 𝟎) 𝑷(𝒌 = 𝟏) 𝑷(𝒌 = 𝟐)

𝑃0,3 1 0.9867 0.96

𝑃1,3 0 0.0023 0.0067

𝑃2,3 0 0.0023 0.0067

𝑃3,3 0 0.00283 0.0083

𝑃4,3 0 0.00283 0.0083

𝑃5,3 0 0.00155 0.00457

𝑃6,3 0 0.00155 0.00457

𝑃7,3 0 6.50E − 06 5.77E − 05

𝑃8,3 0 6.50E − 06 5.77E − 05

𝑃9,3 0 3.55E − 06 3.14E − 05

𝑃10,3 0 3.55E − 06 3.14E − 05

𝑃11,3 0 6.50E − 06 5.77E − 05

𝑃12,3 0 6.50E − 06 5.12E − 05

𝑃12,4 0 0 6.46E − 06

𝑃13,3 0 3.55E − 06 2.79E − 05

𝑃13,4 0 0 3.51E − 06

𝑃14,3 0 3.55E − 06 3.14E − 05

𝑃15,3 0 4.44E − 06 3.95E − 05

𝑃16,3 0 4.44E − 06 3.95E − 05

Page 125: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-109

𝑃17,3 0 4.44E − 06 3.51E − 05

𝑃17,4 0 0 4.40E − 06

𝑃18,3 0 4.44E − 06 3.95E − 05

𝑃19,3 0 1.02E − 08 2.73E − 07

𝑃20,2 0 0 1.02E − 08

𝑃20,3 0 1.02E − 08 2.62E − 07

𝑃21,3 0 1.02E − 08 2.42E − 07

𝑃21,4 0 0 3.06E − 08

𝑃22,3 0 1.02E − 08 2.73E − 07

𝑃23,3 0 1.02E − 08 2.42E − 07

𝑃23,4 0 0 3.04E − 08

𝑃24,3 0 1.02E − 08 2.73E − 07

𝑃25,3 0 1.02E − 08 2.02E − 07

𝑃25,4 0 0 6.08E − 08

𝑃25,5 0 0 1.02E − 08

𝑃26,3 0 1.02E − 08 2.42E − 07

𝑃26,4 0 0 3.04E − 08

Note: Only the non-zero state probabilities are depicted in the table. Of course, as the time step

increases the whole state space will be covered.

For discussion and illustration purpose, consider Table 35 and Table 36.

1. For 𝑘 = 1, the top event probability 𝑃𝑛,5(𝑘∆𝑡) = 0. This implies that the system overflow

scenario will not occur in a single time step given that the system was in the nominal region at

𝑘 = 0. It can be observed that even though the system does not result to a condition, there exist

some probabilities for the system to make transitions from the system nominal operation state

to several other system states. This can be evidently observed in the table, where the 𝑃𝑛,𝑗(1)

elements for 𝑛 ≠ 0 and 𝑗 = 3 have non-zero probabilities. These elements can be interpreted

as the probabilities of the system states transition due to the stochastic nature of the unit

failures, however without leading to any system failure condition. This provide the analyst an

important information that a defined top event in terms of state vector magnitude does not

Page 126: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-110

necessarily occur even if a hardware/component failure may occur at a given time. This is due

to the fact that the system dynamics have been accounted for, and the initial/previous position

of the state vector at ((𝑘 − 1)∆𝑡) in the discretized space and its subsequent evolution for a

given system state dictates the top event probabilities. In contrast to the FT analysis where

instant system failure occurs, or the top event probability is dictated only by the state of the

basic events, thus neglecting dynamic evolution of state vector. The methodology also provides

significant quantitative information on the possible states a system can take at any point in

time. Furthermore, the probability of the system being in normal condition i.e., all units normal

and liquid level at nominal region is 0.986. This value implies the dynamic reliability of the

system or an explicit representation of hardware state and state variable. Besides the normal

operating condition, the most probable state that the system can take in a single time step are

𝑃3,3(1) = 𝑃4,3(1) = 2.84E − 03. 𝑃3,3(1) and 𝑃4,3(1) both implies a change in system state

[(𝑛1 = 0 ∩ 𝑛2 = 1 ∩ 𝑛3 = 0)𝑎𝑛𝑑(𝑛1 = 0 ∩ 𝑛2 = 2 ∩ 𝑛3 = 0)], but constant state variable

magnitude (Region-3).

2. For 𝑘 = 2, summing up all the elements for 𝑗 = 5, we obtain the top event probability of

system overflow is 𝑃𝑛,5(2) = 1.02E − 08 and system drained is 𝑃𝑛,1(2) = 0. And the

probability of the system being in normal condition is 0.96. It is self-explanatory that the model

output explicitly identifies system configuration, state variable magnitude and unit state order.

The obtained results of 𝑃𝑛,𝑗((𝑘 + 1)∆𝑡) can be validated by checking the necessary condition

which is; ∑ ∑ 𝑃𝑛,𝑗(𝑘∆𝑡)𝐽𝑗=1

𝑁𝑛=1 = 1. This validation assures the analyst that the computation was

performed correctly with the computational cells covering the entire space without overlapping,

and that the probability of finding the state vector 𝑥 in the discretized space 𝑗 for a given system

configuration 𝑛 at any given time step 𝑘∆𝑡 is 1.00.

Conclusion and Comparison

In this chapter, a detail analysis of the BS was performed using FT/ET analysis, classical Markov

model, DFM and coupled Markov-CCMT method. It was observed that the dominant classical

techniques (ET/FT) have limitations in the modeling and treatment of time-dependent interactions

that shape dynamic event sequences, and state variable evolution. For instance, for the BS, the end

states depend on the order of unit failure, timing of failure events and magnitude of the state

Page 127: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-111

variable. Classical PRA approach which is static in nature do not have the capability to capture

these interactions and may omit risk-significant event sequence. Classical Markov model have the

capability to capture time-dependent system behavior through state transition matrix. However,

the model lacks the capability to capture system dynamics. Dynamic PRA methods attempt to

address the above-mentioned issues and drawbacks by accounting simultaneously for the time

element, stochastic state transition and dynamic system evolution. A comparison of the

methodologies is presented below.

1. Predicted benchmark system failure probability

The failure probability of the benchmark system predicted by each of the methodologies for four

(4) time steps are presented in the tables below.

Table 37: Predicted failure probability of the BS using FT (binary)

FT (binary) Predicted system failure probability

𝒕 = 𝟏 𝒕 = 𝟐 𝒕 = 𝟑 𝒕 = 𝟒

Total failure probability 3.15𝐸 − 03 6.36𝐸 − 03 9.61𝐸 − 03 1.29𝐸 − 02

Table 38: Predicted failure probability of the BS using FT (multi-state)

FT (multi-states) Predicted system failure probability

𝒕 = 𝟏 𝒕 = 𝟐 𝒕 = 𝟑 𝒕 = 𝟒

Overflow 1.46𝐸 − 05 5.82𝐸 − 05 1.31𝐸 − 04 2.33𝐸 − 04

Drained 1.02𝐸 − 08 8.16𝐸 − 08 2.75𝐸 − 07 6.53𝐸 − 07

Total failure probability 1.46𝐸 − 05 5.83𝐸 − 05 1.31𝐸 − 04 2.34𝐸 − 04

Table 39: Predicted failure probability of the BS using ET

Event Tree Predicted system failure probability

𝒕 = 𝟏 𝒕 = 𝟐 𝒕 = 𝟑 𝒕 = 𝟒

Overflow 2.91𝐸 − 05 1.83𝐸 − 04 5.69𝐸 − 05 1.29𝐸 − 03

Drained 3.06𝐸 − 08 2.45𝐸 − 07 8.26𝐸 − 07 1.96𝐸 − 06

Total failure probability 2.91𝐸 − 05 1.83𝐸 − 04 5.70𝐸 − 04 1.29𝐸 − 03

Page 128: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-112

Table 40: Predicted failure probability of the BS using Markov model with the qualitative

consideration taken in Table 12.

Markov model Predicted system failure probability

𝒕 = 𝟏 𝒕 = 𝟐 𝒕 = 𝟑 𝒕 = 𝟒

Overflow 0 2.91𝐸 − 05 8.68𝐸 − 05 1.73𝐸 − 04

Drained 0 0 0 6.12𝐸 − 08

Total failure probability 0 2.91𝐸 − 05 8.68𝐸 − 05 1.73𝐸 − 04

For the dynamic methods presented in the below tables, the initial condition of the tank level was

assumed to be in the nominal region.

Table 41: Predicted failure probability of the BS using Markov-CCMT model

Markov-CCMT model Predicted system failure probability

𝒕 = 𝟏 𝒕 = 𝟐 𝒕 = 𝟑 𝒕 = 𝟒

Overflow 0 1.02E − 08 9.17E − 05 5.11𝐸 − 04

Drained 0 0 1.86E − 07 2.15E − 06

Total failure probability 0 1.02E − 08 9.19E − 05 5.14E − 04

Table 42: Predicted failure probability of the BS using DFM

DFM Predicted system failure probability

𝒕 = 𝟏 𝒕 = 𝟐 𝒕 = 𝟑 𝒕 = 𝟒

Overflow 0 3.345E − 05 9.54E − 05 1.86E − 04

Drained 0 1.36E − 07 4.985E − 06 2.345E − 05

Total failure probability 0 3.36E − 05 1.00E − 04 2.09E − 04

2. Multistate Modelling

Fault Tree Analysis:

Classical FTA is based on binary representation of component/system states, and hence only two

states (normal and failed) are analyzed for system reliability assessment. However, realistically,

there is always a possibility for a component/system to have several failure modes. This is

especially true while modelling a redundant system. For instance, a valve can fail-high, fail-closed,

Page 129: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-113

fail in 50% position, etc.; and a system can overflow, drained or quasi-stable. In FTA, component

failure modes are treated independently, rather than in an integrate fashion (refer Appendix I). This

result in an overestimation of system failure probability. Furthermore, for a system with multiple

failure modes, each top event requires a separate construction of FT.

Event Tree Analysis:

Similar to the FTA, classical ETA is also based on binary representation of component/system

states. ETA being an inductive approach provide a possibility to model multiple component or

system states, in the sense that every possible failure mode of a component results in multiple

system states. It may be observed from the BS analysis that, each unit failure mode results in 9

possible system states. However similar to the FTA, all failure modes of a component are treated

independently, and each failure modes requires a separate analysis and construction of ET. Hence

for a component with multiple failure modes the analysis become tedious and time consuming.

Classical Markov Model:

The classical Markov model clearly provide a superior way to analyze systems or components

with multiple states as compared to the classical FT/ET (Appendix I). Markov model of a

component/system takes into account the competition among states of a component, i.e., the

likelihood of occurrence of a particular component state among all possible states. This provide

the analyst a realistic result and does not overestimate the system failure probability. At a system

level, it may be observed from the BS analysis that the model results in 27 distinct system states.

Dynamic Flowgraph Method:

DFM is based on a multi-valued logic, where multiple component/system states are expressed in

terms of causal relationships among system parameters. Any number of component/system states

can be enumerated in the decision table. For instance, a three-state variable (fail low, fail high and

fail at 50% position) is represented as three inputs to a decision table resulting in different output

of the decision table. Furthermore, a single DFM model once developed can analyze any system

condition of possible interest, i.e., any number of top event of interest can be evaluated with a

single model. This is a significant advantage over the classical PSA techniques.

Coupled Markov-CCMT Model:

Markov-CCMT model is based on a multi-state system characterization, with the feature to

represent an arbitrary number of system states depending on the number of component states and

Page 130: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-114

partitioned state variable states. It can be observed from the analysis of benchmark system that the

model result in a 135 distinct system states. This provide significant information to an analyst with

regard to possible states a system can take. However, a drawback in explicit states definition is

that it is difficult to visualize or foresee the total set of possible states a system can take prior to

scenario development.

3. Time Dependencies

Fault Tree and Event Tree Analysis:

The classical FT/ ET is based on static logic modelling, and hence do not account for the time

element (exact timing of failure event) and lacks the capability to accurately quantify the event

sequence probability deviation with time besides the state variable evolution in time.

Classical Markov Model:

The Markov model is a rigorous time-dependent model with explicit modelling of time element

during states transition. Time-dependent system state probabilities can be obtained once the model

is constructed. However, an accident sequence with state variable evolution in time cannot be

captured in the model due to the fact that system dynamics is ignored in the model.

Dynamic Flowgraph Method:

The DFM can explicitly represent time element as well as the evolution of the state variable in

time besides depicting the type of unit failure. All the PIs generated via DFM model are time-

stamped and dictates the type of system failure scenario. The backtracking process can also be

performed to unveil system state in the previous time steps (dependent on the depth of

backtracking), as was demonstrated in this chapter. This allow one to observe the development of

event sequence in time.

Coupled Markov-CCMT Model:

The coupled Markov-CCMT model can explicitly account for the time element as well as the state

variable evolution. Besides the unit states, the exact timing of failure event is taken into account

that eventually influence the end state (overflow or drained). The methodology utilizes real time,

and thus emphasize on the exact timing of unit failure events as compared to the DFM model.

Page 131: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-115

4. System Dynamics

Fault and Event Tree Analysis:

The classical FTA and ETA represent a system state with a set of success/failure states of the

system components. The methodologies are neither developed nor is deliberated to model system

dynamic response and its evolution. Thus, the classical methodologies lack the capability to

capture unit/component dynamic interactions. It was observed that ET technique provide a better

approach compared to FT to model systems that response dynamically to IEs, in that, it can provide

a correct failure logic and a qualitative assessment of system end states.

Markov Model:

The classical Markov model is intended for discrete time discrete system state representation, but

not to model system dynamics or state variable evolution, i.e., the deterministic aspect. The model

is oriented to system hardware states modelling rather than system dynamics. Hence it lacks the

capability to model dynamic system evolution.

Dynamic Flowgraph Method:

DFM provide a favorable approach to model system dynamics by explicit representation of time

element and state variable evolution. Dynamic system behaviors are represented as a series of

discrete state discrete time transitions and is modelled in the transition box decision tables.

Furthermore, dynamic consistency rules are applied to eliminate impossible events and generate a

consistent accident sequence. However, decision tables have to be constructed separately for

varying time steps size and state variable intervals. The methodology is oriented towards and

sensitive to the partitioned state variable interval size. A drawback of the methodology is that, the

influence of state variable evolution on the failure/demand rate of the units cannot be accounted.

The methodology rather focuses on probabilistic system dynamic and provide an overview of the

evolution of failure as well as normal events for each time steps. Overall DFM can accurately

represent, analyze and uncover all risk-significant dynamic event sequence.

Coupled Markov-CCMT model:

Due to the limitation of the classical Markov model to account for system dynamics, CCMT is

coupled with Markov model to represent system dynamics in discrete-space discrete-time. The

joint transition probability matrix of the model explicitly depicts evolution of the state variable,

the conditional dependencies between the unit states. Thus, the model captures the unit states

Page 132: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-116

transition caused by deterministic laws as well as by stochastic nature of the units. This feature of

the methodology enables one to simultaneously account for system dynamics and stochastic state

transition (tightly coupled). In addition to accounting for time-dependent system states probability,

the model has the capability to capture the possibility of failure frequency deviation with state

variable evolution i.e., likelihood of a subsequent failure given an IE due to the state variable

evolution. For instance, if all the units are in normal state and the state variable make a transition

from region 𝑗′ → 𝑗 = 2 → 3 at 𝑡 = 0, the deterministic law will demand unit-1 to remain open,

unit-2 to turn off and unit-3 to open. Hence, the possible unit failure states are: unit-1 can fail-

closed, unit-2 can fail-open and unit-3 can fail-closed. Unit-1 fail-open failure rate is not taken into

account since this state of the unit does not contribute to the system failure conditioned upon that

the control laws demand the unit to open. Hence, for this particular case unit-1 fail-open rather

contribute to the success of the system at this particular time step and state variable magnitude

transition. The same argument holds for the case of unit-2 and unit-3. This dynamic scenario and

system response to an event thus changes the probability of unit failure, besides the deviation of

unit failure rate with time.

5. Event Ordering

Fault Tree Analysis:

The FTA is not intended to model state or failure event ordering, rather it represents the top event

in terms of combinations of basic events (minimal cut-sets) without any particular ordering index.

Event Tree Analysis:

In ETA, the sequence ordering in pre-set by the analyst as was observed and pointed out in the ET

of the benchmark system. Hence, the methodology lacks the capability to capture risk significant

dynamic event sequences that can probabilistically evolve from dynamic system interactions,

which would then remain uncovered.

Classical Markov Model:

In the Markov model, the order in which a unit/component state occur in the accident sequence is

explicitly modelled due to the dependence of top events on state ordering. This is especially true

for systems with multiple top events. The state sequence and possible number of system state is a

priori fixed by the analyst. It can be observed from the BS analysis that each of the 27 states have

Page 133: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-117

unique state ordering, and the probability of occurrence of a state is dictated not only by time and

transition rate, but also by the order in which unit failure occurs. The state ordering is simplified

by categorizing the system states in a layer wise approach (as done in this case).

Dynamic Flowgraph Method:

In DFM, the event sequence is not pre-determined by the analyst but rather the sequence evolves

probabilistically from the time-dependent system model, and dynamic interactions among the

units. The methodology strictly takes into account the order of unit failures in each time step. This

allow one to observe the time-dependent scenario development in an orderly fashion. The dynamic

accident sequence is presented as a set of timed-prime implicants for any top event of interest with

explicit representation of the state variable magnitude and time element. Size of the PIs as well as

the number of PIs increases with increasing time steps, i.e., similar to the logic ‘AND’ and ‘OR’

gate in FT where AND gate increases the MCS size and OR gate increases the number of MCSs.

Coupled Markov-CCMT:

Similar to the classical Markov model, state sequence is explicitly modelled, and the possible

number of system states are a priori set by the analyst. This is also a limitation since it is difficult

to envision all the possible system states prior to scenario development. Only the state probabilities

changes, and the model is memoryless, in that, time-dependent sequence development cannot be

observed. Unlike DFM, the order, size and number of system states in Markov-CCMT model

remains fixed throughout the analysis. Here, DFM provide an advantage as it allow one to observe

the dynamic event sequence.

6. Computational Demand, Complexity and Time

Fault Tree and Event Tree Analysis:

The classical FTA and ETA are well-established methodologies that can be implemented

systematically requiring less computational resource and time. The time requirement however may

increase significantly depending on the number of top events and IEs for FT and ET respectively.

Classical Markov Model:

For system with large number of possible system states, the computational demand and complexity

increases significantly. Furthermore, construction and modelling of the transition diagram can be

considerably time consuming depending on the number of states.

Page 134: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-118

Dynamic Flowgraph Method:

In comparison with other methodologies, DFM provide the flexibility for system analysis via

inductive as well as deductive approach. However, construction of decision tables can become

complex and time consuming with increased number of system parameters and discretized states,

requiring other pseudo codes for decision table construction. For instance, decision table TT1 for

the BS was constructed in MATLAB. The methodology also requires a highly time-dependent

system model which may require significant amount for time for model development. The results

obtained from DFM model requires significant post-processing so that they can be integrated into

an existing classical PRA model. The post-processing via backtracking process, and subsequent

generation of timed-FTs requires significant amount of time and diligent application of consistency

rules in every branching point, thus requiring a high analytical skill level.

Coupled Markov-CCMT:

It was observed from the BS analysis that the computational demand for coupling physical and

stochastic model can be very intensive, complex and timing consuming. The computational

increases significantly with the number of state variables, possible system states and discretized

number of computational cells in the state-space. This phenomenon is well known as state space

explosion. The size of the transition matrix can become very large requiring significant amount of

time for its evaluation. In contrast to the above, the methodology provides features to validate the

model at any point in time during its implementation. Another drawback of the methodology is

that, it requires a highly time-dependent system model. Furthermore, post-processing of the results

obtained from Markov-CCMT model requires significant time. Last but not the least, the analytical

skill level required to model and implement the methodology is considerably demanding.

7. Interface

Fault Tree and Event Tree Analysis:

FTA and ETA are user friendly, and there exist several well-established computer codes such as

CAFTA, FaulTree+ and RiskSpectrum.

Classical Markov Model:

Solving the set of coupled ODEs can be challenging with increased number of system states and

transition among these states. However, there exist many well established numerical methods for

Page 135: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-119

obtaining the ODE solutions and can be systematically implemented in computer codes such as

Fortran95, C and Python.

Dynamic Flowgraph Method:

System modelling and analysis in DFM can be performed in DYMONDA code, a graphical user

interface. This is a significant advantage over other dynamic techniques which does not have an

integrated and user friendly interface.

Coupled Markov-CCMT Model:

There exist no well-established integrated computer codes for implementation of the methodology.

Stochastic nature of the components and the deterministic aspects of the system have to be

modelled separately using different codes, and then eventually coupling the two model. For the

BS, the Markov model is implemented in Fortran95 and CCMT or deterministic system model is

implemented in MATLAB. Thus, coupling of Markov-CCMT model is relatively complex and

demands high computational time especially for realistic systems.

Page 136: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-120

CHAPTER 5: AN INTEGRATED APPROACH FOR RELIABILITY ASSESSMENT

OF PASSIVE SAFETY SYSTEM

Introduction

This chapter presents a preliminary roadmap and activities to be carried out for the project on

“Design and Performance Assessment of Passive Engineered Safety Features in Advanced Small

Modular Reactors”. the method of selection of the technology and system to be analyzed is

presented in Section 5.2. Section 5.3 discusses a novel approach to modelling generic passive

safety systems (PSSs) in an integrated framework for dynamic reliability assessment. Section 5.4

provide a detail description of a generic isolation condenser system (ICS) to be used as a

benchmark system for the CRP. Note that the current research is not design/plant specific but

concerns a typical design of an ICS implemented in iPWR SMRs. Section 5.5 characterize the ICS

into macro-components or units, with each unit modelled separately using FT (failure rate

evaluation method) and Markov model. Section 5.6 integrates the individual units to predict the

overall ICS system behavior. Note that only sample results are presented for illustration purpose.

Research Project Roadmap

This sub-section provide a roadmap for the coordinated research project (CRP) on, “Design and

Performance Assessment of Passive Engineered Safety Features in integral PWR-type Small

Modular Reactors”. The task and responsibilities include:

• For 2017: literature survey of defence-in-depth being adopted in SMR designs and draft

a technical report on technical description of passive engineered safety features in SMRs;

• For 2018: develop manual and support CRP participants in using certain PSA techniques;

• For 2019: draft the scope and main content of a technical document (project outcome).

Recognizing and taking into account the above responsibilities to be fulfilled, a review of current

SMR technology and passive safety systems in advanced reactor designs was performed as an

initial step of the project. An overview of the project roadmap is presented in Figure 24.

Page 137: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-121

Figure 24: Proposed coordinated research project roadmap

Aforementioned in Chapter 4, Markov-CCMT model was chosen over DFM to model passive

systems due to its capability to capture time-dependent dynamic interactions. The author intends

to provide the basis of how dynamic PRA can be more applicable and suitable for passive systems

analysis in iPWR-type SMRs.

1. SMRs are simple in design; thus, fewer number of components that address both steady

and transient conditions. Yet even for smaller core, off normal reactivity control must be

as quick as large reactors. Thus, post shutdown system must remove decay heat with little

human intervention for approximately 72 hours mission time. Thus, the extent to which

dynamic analysis like that presented in Chapter 4 should be investigated in order to assure

that classical PRA is sufficient or insufficient;

2. Provided that iPWR-type SMRs as designed in simple terms with less number of

components, the phenomenon of state-space explosion can be investigated to develop and

demonstrate dynamic PRA methods.

3. Passive systems performance is not only dependent on the hardware state, but also on

critical system parameters such as presence of non-condensable, heat loss, fouling, etc.

These parameters which influence the system dynamics and hence the functional reliability

must be taken considered in the analysis. The dynamic methods presented in Chapter 4

2017 2018 2019 2020

2016

• Research proposal awarded

• Review of SMR technology

• Review of Passive Safety Systems

• Review of Passive Safety Systems in iPWR SMRs

• Selection of Benchmark PSS to be analyzed

• Selection of methodology

• FMEA• Development of

Stochastic and Physical model

• Classical approach to system modelling

• Coupling of stochastic-physical model

• Reference case simulation

• Initial Drafting of IAEA-UOIT project

• Integration of draft from all participating countries

• Final drafting of the Technical Document

17th October:

Research Proposal

18th July

Page 138: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-122

have the capabilities to account for these parameters in an integrated fashion. However, the

applicability of the techniques to model and analyze passive systems reliability should be

demonstrated via a benchmark system.

The IAEA defines PSSs as “a system that is composed entirely of passive components and

structures or a system, which uses active components in a very limited way to initiate subsequent

passive operation”. PSSs can be categorized into four (4) class: [IAEA-TECDOC-626]

I. Category A: Characterized by systems that has no signal inputs of intelligence, no

external power sources or forces, no moving mechanical parts and no moving working

fluid, e.g., accumulators.

II. Category B: Characterized by systems that has no signal inputs of intelligence, no

external power sources or forces, no moving mechanical parts but have a moving

working fluid, e.g., passive containment cooling systems.

III. Category C: Characterized by systems that has no signal inputs of intelligence, no

external power sources or forces, but composed of moving mechanical parts with or

without moving working fluids, e.g., relief valves.

IV. Category D: Intermediary zone between active and passive where the execution of the

safety function is made through passive methods, i.e., passive execution/active

initiation, e.g., emergency shutdown systems based on gravity.

A comprehensive review of active, passive and hybrid safety systems implemented in SMRs

technology with focus on iPWR design was performed. Additionally, the author categorized the

PSSs based on their functions, including:

1. Passive residual heat removal systems (e.g., mPower)

2. Passive safety injection system (e.g., SMART)

3. Passive containment cooling system (e.g., NuScale)

4. Passive/automatic depressurization system (e.g., CAREM).

The iPWR SMR designs that implements passive systems to achieve the above mentioned

functions are provided as an example. A stepwise assessment and flowchart of the reviews

performed to select the reactor design and passive safety system to be used as a benchmark for the

project is depicted in Figure 25.

Page 139: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-123

Figure 25: Stepwise review and flowchart for design and system selection

The passive ICS belonging to the category of passive reactor depressurization systems was selected

to be used as a benchmark system for the project. According to the IAEA classification on passive

systems, the ICS is within the Category D. The basis for the selection of ICS as a benchmark

system includes:

• The ability to maintain reactor coolant system pressure without the loss of primary coolant;

• The ICS has been implemented in iPWR SMRs (e.g., CAREM, IRIS, AHWR) with varying

objectives such as reactor depressurization, to maintain reactor hot standby and to reduce

the frequency of SRVs operation;

• The ICS has been used as a benchmark system for validation of developing methodologies

for reliability assessment of passive safety systems [IAEA-TECDOC-1752];

• Large reactor designs such as ESBWR has implemented ICS. This increases the availability

of research materials on ICS, physical modelling approach and experimental data;

• An ongoing effort within the nuclear community to model and predict system behavior

based on natural circulation.

After the selection of reactor design (i.e., iPWR) and passive safety system to be analyzed (i.e.,

the ICS), the next step is the selection of a methodology to assess and predict the deterministic as

well as probabilistic behavior of the isolation condenser system. A comprehensive review of

current methodologies for assessment of PSSs was performed, of which reliability methods for

Passive Safety Systems in NPPs

• Review of passive safety

systems

• Categorization of PSSs by class

and safety functions

• PSSs in SMR Technology

Safety Systems in NPPs

• Active safety systems

• Passive safety systems

• Hybrid safety systems

SMR Technology

• Concept of modularity

• Concept of integrated

reactor coolant system

• Multi-module concept

SMR Technology Development

• Integrated PWR design

• BWR reactor design

• Heavy water reactor design

• Integrated molten salt reactor

Selected Technology and System

• Integrated PWR design

• Passive Isolation

Condenser System

Page 140: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-124

passive safety functions (RMPS) [Ricotti et al. (2002); Jafari et al. (2003); Marques et al. (2005)]

and Assessment of Passive System Reliability (APSRA) [Nayak et al. (2007, 2008(a))]

methodologies provide a promising approach [IAEA-TECDOC-1752 (2014); Zio et al. (2009)].

The two methodologies differ from the fact that APSRA attributes the passive system failure to

hardware failures (e.g., condensate return valves), whereas RMPS methodology emphasize more

on the uncertainties arising from physical phenomenon such as heat loss, non-condensable

fraction, oxidation, etc. In PSSs system dynamics have a huge impact on its behaviour. Both the

methodologies attribute the passive system failure to the deviation of system parameters and

component failures, however the two are treated separately, i.e., independent modelling of system

deterministic and probabilistic aspect. However, the performance of passive systems can be

significantly affected by the system dynamics especially due to the low driving force. For instance,

during the course of passive system operation, a component can fail due to its stochastic nature

which will in turn affect the system parameters. This change in system parameters can further

influence the state of the hardware components. These complex dynamic interactions can rapidly

result to a system failure and may evolve to a system state not anticipated or predicted by the

deterministic assessment and classical PSA approach. To the author’s knowledge, the current

passive system reliability assessment methodologies do not account for these complex dynamic

interactions and subsequent probabilistic dynamic system evolution. This motivates the author to

develop a novel approach and provide an integrated framework to account for risk-significant

scenarios arising from dynamic interactions. The integrated dynamic approach is particularly

important for passive safety systems assessment that operates extensively on natural and physical

laws that have a small driving force, as compared to the active systems that operates on high

driving forces from external input energy. This implies that a small deviation in critical system

parameters can significantly influence the system performance and hence the reliability. Given an

IE, it is very likely that the state variables can make a transition out of the control space very

quickly as compared to active systems. Furthermore, there exist uncertainties with regard to

complex physical phenomenon such as natural circulation, thermal stratification, etc., which are

not well understood or modelled till date. The proposed methodology has the potential to account

for these complex interactions, capture the physical phenomenon by accounting for state variable

evolution, and provide a realistic estimate of failure probability associated with passive systems.

Page 141: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-125

An Integrated Framework for Dynamic Reliability Assessment of Passive Safety

Systems

A novel approach for systematic modelling and integrated analysis of passive safety systems

(PSSs) is outlined in this sub-section. The approach provides a framework to couple the system

deterministic evolution with the system stochastic nature. The proposed approach consists of

several steps and provide a roadmap for the IAEA-UOIT CRP as well as to perform a meaningful

comparison between the classical PSA techniques and dynamic methodologies. The objectives of

the methodology include:

• Provide a coherent approach to modelling PSSs in NPPs;

• An integrated framework to couple deterministic and probabilistic aspect of a system;

• A cohesive treatment of both aleatory and epistemic uncertainties;

• The inclusion of PSSs as a frontline system in accident sequence analysis.

Note that the methodology in many ways is explained using the passive ICS as an example so that

the readers could comprehend the approach with ease. The project roadmap and the proposed

methodology flowchart is shown in Figure 26.

1. System Identification and description

The first step of the proposed methodology involves the identification and selection of the passive

safety system to be analyzed. A brief description of the system and principle of operation may be

provided along with the scenarios in which the system is expected to operate.

2. Accident/transient scenario

The descriptions of scenarios (transient/accident) for which the passive safety system is designed,

along with the system state during normal reactor operations. A brief operational characteristic

and the system initiation following a reactor transition from normal to transient condition.

3. Definition of System Mission

The mission of a system are the objectives to be accomplished by the system under a priori defined

normal or accident scenarios. Generally, the sets of objectives are predefined by the designers and

can be correlated. Overall, the scenario and time at which the system is required to operate as well

Page 142: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-126

as the duration of operation (mission time) must be manifested in this step. For example, the ICS

maintains the RCS pressure by removing decay heat from the reactor for a period of 72 hours.

4. Success and Failure Criteria of the System

Success and failure criteria of the system must be manifested for all possible modes of operation

and scenarios. The criteria may include hardware failures, state variables exceeding a pre-defined

threshold limit or activation/operation of other safety systems. For example, the ICS can be

considered failed, if the safety relief valves operate or if the system pressure is above 7.5 MPa.

Often, these criteria are related to the system mission, and hence while defining the criteria, system

mission must be taken into account.

5. Operational characteristics and parameters identification

The operational characteristics and the parameters influencing the operation of PSSs must be

identified. These two factors are correlated since system parameters effect the PSSs operational

behavior. The goal of this step is to comprehend the operating principle and characteristics from a

qualitative view point, and not to accurately model and predict the system behavior. Of course,

this will naturally involve identification of system operating parameters. For example, a natural

circulation operates on coolant density difference between the heat source and sink, and the amount

of heat removed can be dependent on the coolant flowrate, pressure and temperature in the sink

and source side, etc. All the system parameters involved must be identified and listed in this step.

6. Identification of failure modes affecting the system performance

This step involves systematic identification of failure modes affecting the system performance.

This can be achieved by the well-established and commonly used standard techniques such as

failure mode and effect analysis (FMEA), and hazard and operability analysis (HAZOP). The

traditional approach is mostly oriented towards hardware failure modes, however physical (virtual)

phenomenon failure modes that degrade the physical mechanism must be taken into account in

case of PSSs analysis due to its reliance on small driving force. For instance, performance of the

ICS is significantly affected by pipe fouling, oxidations, presence of non-condensable gases,

corrosion, degraded heat transfer, etc. Furthermore, the modes of failure should be prioritized with

regard to the system performance.

Page 143: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-127

Figure 26: The project overview and methodology flowchart

Quantitative Reliability

Evaluation (PDF, CDF)

Coupling of the stochastic-

deterministic model

State space partitioning

(computational cells)

Classical PSA approach

• Reliability assessment

• Importance/sensitivity

analysis

• Uncertainty analysis

System Modelling

• Selection of best estimate code

• Development of model

• Reference case simulation

• Validation of model

Preparation of failure

data base

Identification of

failure modes

Operational characteristics and

system parameter identification

System

Identification

• Definition of system mission

• Success/failure criteria of the system

Identification/screening

of critical parameters

Stochastic system model

(system state-space)

Computation of cell-to-

cell transition probability Computation of stochastic-

dynamic dependencies

Definition of Top Events

Comparison

Sensitivity and

Integrated

Uncertainty analysis

Page 144: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-128

7. Preparation of failure data base

A data base must be developed for the purpose of quantitative assessment. The failure data can be

given in terms of probability of failure or failure rate with an exposure time, representing an

unavailability of a component. Each of the failure modes identified in step 5 must be assigned with

a failure probability or failure rate. Failure data may be obtained from international reliability data

base such as the IAEA, USNRC and IREP.

8. Identification/screening of critical parameters affecting the system performance

This step is a continuation of step 5, in that, all the identified parameters are ranked or screened

according to their importance or the level of influence on the system performance. The critical

parameters are direct indicator of the system performance and include both hardware state as well

as state variables. For example, the ICS operation is significantly affected by drain valve state, the

presence of non-condensable gases, water temperature in the pool, differential pressure, etc.

Importance analysis can be performed in the fault tree analysis to rank the critical hardware

components. However, a physical model is required to rank the physical system parameters. This

step enables an analyst to deduce the system modelling and analysis to a feasible and manageable

state by eliminating the insignificant parameters.

9. Identification of parameters relationship and dependencies

The dependencies between critical system parameters must be identified and taken into account

adequately in the process of quantification. Conditional probability can be used for the classical

approach whereas dependencies arising due to dynamic interactions is treated in the process of

coupling the system stochastic model with system dynamic model. The negligence of relevant

dependencies can result in an inaccurate assessment of system reliability. For example, in the ICS,

operation of the drain valve is dependent of the loop pressure. Loop pressure could be affected by

the heat transfer rate, leading to a reduced heat transfer to the pool and hence further increasing

the pressure in the loop. This could further lead to a reduced opening of the drain valve, resulting

to a rapid increase of failure.

10. Reliability assessment via classical PSA techniques

Once the failure modes are identified, prioritized and failure rate data base are created, system

reliability assessment (qualitative and quantitative) can be performed using the well-known

Page 145: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-129

standard classical fault tree technique. The combination of failure modes resulting to a system pre-

defined failure i.e., the MCSs and its subsequent quantification can be performed using well-

established computer codes such as CAFTA. There can be multiple top events with respect to PSSs

initiation and operation as shown in Figure 27. 𝑃𝑆𝑆𝑆 implies passive safety system start-up, ��𝑆𝑆𝑆

implies PSS failure to start, 𝑃𝑆𝑆𝑂 implies PSS operation, ��𝑆𝑆𝑂 implies failed operation, 𝑃𝑆𝑆𝑃 implies

PSS successful performance or it meet the design criteria and ��𝑆𝑆𝑃 implies passive safety in

operation but failed to meet the criteria. For example, PSSs based on natural circulation in principle

will not fail as long as there is a heat sink and source. However, it may not meet the success criteria

such as heat transferred, coolant flowrate, etc.

Figure 27: Representation of passive safety system in classical ET/FT

11. Stochastic modelling of the system

This step involves modelling of the system using the well-known time-dependent Markov chain.

The system is described by a finite number of system states, and the probability of the system

being in any of the pre-defined states is computed. Each of the system states are further

characterized by the individual component states, i.e., the discretized state-space is characterized

by the smallest element in the system (basic event). The ordering of states can also be performed

if it effects the end state of the system. To reduce the number of system states and avoid state space

explosion, the system can be characterized into macro-components or units, where the individual

units can be modelled separately, e.g., using fault tree.

𝑃𝐼𝐸

Passive safety

system start-up

IE Passive safety

system operation

Passive safety system

performance

��𝑆𝑆𝑆

𝑃𝑆𝑆𝑆

𝑃𝑆𝑆𝑂

��𝑆𝑆𝑂

𝑃𝑆𝑆𝑃

��𝑆𝑆𝑃

Fault Tree Fault Tree Fault Tree

Page 146: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-130

12. Development of system physical model

This step involves a detail development of the system physical model using qualified best estimate

thermo-hydraulics system codes (e.g., RELAP5) or a simpler standalone model. Uncertainties arise

in the modelling process especially due to lack of knowledge of physical phenomenon in PSSs,

lack of operating experience, the input variables, approximation in system geometry, etc. This in

turn effects the predicted system behavior. Uncertainties in the predicted system behavior can be

reduced using or by comparing against experimental data in the modelling of physical

phenomenon.

13. Model Validation: reference or design case system simulation

Once the physical model is developed, a standard reference case can be run to validate the system

model. Select an initiating event that requires operation of the PSS and observe the system behavior

and performance. For instance, for the ICS, a closure of MSIV or reactor scram can be selected as

an initiating event that increases the system pressure and eventually demands the ICS to operate.

The system model validation can also be performed if experimental data are available.

14. Define the top event of interest

In contrast to the classical approach (FT) and Markov model, the top events in dynamic

methodologies are defined in terms of the state variables. For instance, for the ICS, the top event

can be defined as ‘RCS pressure above a defined threshold limit’. Of course, the computational

complexity and time increases significantly with the increase in state variables. Again, taking the

ICS as an example, the top event can be described as ‘system pressure and peak cladding

temperature above the threshold limit’. As an initial step and with the intention to demonstrate the

capabilities of the methodology in modelling PSSs, only one state variable is considered, i.e., the

RCS pressure.

15. State-space discretization: controlled state variables

The state space is discretized into finite number of disjoint computational cells covering the entire

space, with the control region defined by the analyst. The control space can be identified with the

knowledge of system set-points and boundaries. Information of the top event defined in step 13

can be utilized for the purpose of state-space discretization and vice versa. Step 13 and 14 are in a

sense complementary to each other. The number of computational cells in the state-space and

Page 147: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-131

control space is user defined. Depending on the analyst choice, the state space can be single or

multi-dimensional. For example, a 3-dimensional sate space can consist of system pressure, peak

cladding temperature and reactor water level. Of course, the computational complexity will

increase significantly with the increase in the number of computational cells and state variables.

On the other hand, too less number of cells will result in an inaccurate prediction of the system

behavior. Hence, in the discretization process, the analysts should keep in mind a balance between

the prediction accuracy and computational complexity have to be made.

16. Computation of cell-to-cell transition probability

The system dynamics is modelled as transition between the discretized computational cells in the

state space, i.e., the probability of making a transition from one cell to the other. Here, one is

interested in the evolution of the state variables in the state-space for a given system state. The

discretized system state defined in step 10 is an input, whereas the state variable trajectories is the

output. The number of simulation run is dependent on the number of distinct system states, which

can be very large and can quickly become unmanageable. Depending on the knowledge of the

system, state merging can be done to reduce the number of runs. The computation of cell-to-cell

transition probabilities are explained in detail in chapter 3 and 4.

17. Stochastic-dynamic dependencies model

This step is arguably the most important part of the methodology. The stochastic time-dependent

system model in step 10 is modified taking into account dynamic interactions between the system

hardware and state variables. This step captures the dependencies between the two elements as

well as dependencies arising from human error. For example, the system pressure will deviate

based on the state of the drain valve, or the demand frequency of the drain valve can increase or

decrease in a certain time interval depending on the system pressure. These dynamic interactions

can affect the failure rate magnitude of a component. The term “stochastic-dynamic dependencies

model” is used due to the fact that the model simultaneously takes into account the random

component failures as well as dependencies arising from system dynamic evolution.

18. Coupling of the stochastic and deterministic model

The cell-to-cell and system state transition probabilities obtained from step 13 and 14 respectively

are merged into a single state transition probability matrix that describes the system stochastic and

Page 148: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-132

dynamic behavior. The system dynamic evolution is transformed into a Markovian model, where

the probabilistic mapping is performed via the state transition matrix. A detail theoretical and

numerical implementation of the coupling process is given in chapter 3.

19. Quantitative Reliability Evaluation

This step involves a systematic computation of state probabilities at any given point in time

(similar to the classical discrete Markov chain). Once computation of state probabilities is obtained

for each user defined time step, the probability density function (PDF) and cumulative distribution

function (CDF) of a predefined top event can be determined. The statistical importance of any

system states/configuration can also be computed for a given top event.

20. Assignment of probability distributions to critical parameters

In order to add credit to the point estimate predicted system behavior, critical parameters identified

in Step-7 can be assigned with some probability distributions with the objective to perform an

uncertainty analysis. The choice of the distributions is dependent on the state of knowledge of the

parameter, availability of data and expert judgement. If there is a very limited knowledge of the

system parameter, a uniform (Gaussian) distribution can be assigned i.e., all the data points within

the bounded limits are equally likely. The assignment of distributions to the parameters must be

done with great care since it significantly affects the predicted reliability of the PSS.

21. Integrated uncertainty evaluation

Upon assignment of probability distributions to the critical parameters which include both

hardware and physical parameters, propagation of the distribution in the model can be performed

using standard techniques such as direct Monte Carlo simulation. The superiority of the

methodology is that, both the aleatory and epistemic uncertainties can be captured in an integrated

fashion. Uncertainties arising due to lack of knowledge of physical phenomenon can be accounted

in the probabilistic state variables mapping, whereas uncertainties due to the random component

failure is accounted in the stochastic model.

The Benchmark Passive Isolation Condenser System

1. System identification and description

Page 149: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-133

The isolation condenser system (ICS) is employed in many advanced generation and innovative

reactor designs (Gen III and Gen III+), including integrated pressurized water reactor (iPWR) type

small modular reactors (e.g., CAREM25, IRIS). The ICS consist of a heat exchanger, IC pool,

isolation valves, drain valves, bypass drain valves and vent valves. The schematic diagram is

depicted in Figure 28. The principle of operation is based on natural circulation.

Figure 28: Passive Isolation Condenser system

The ICS operates in closed loop natural circulation mode (gravity driven) by condensing the

incoming saturated steam inside the IC HX and returning the condensate back via a dedicated

condensate return line to the RPV. This operational mechanism of buoyancy driven pump is

created by the heat source and sink with an elevation difference between the RPV and the IC water

pool. Decay heat removal from the reactor (heat source) is achieved by transferring heat to the

isolation condenser (IC) water pool (heat sink) via IC heat exchangers (HX) that are immersed in

the IC pool, located outside the containment and above the RPV (See Figure 28). Steam in tube

side of the IC HX is condensed by boiling pool water in the shell side of the HX and venting the

evaporated pool water to atmosphere. Due to the opening of the valve on the condensate line in

order to trigger its operation, the ICS comply with the IAEA Category D passive system, which

addresses the intermediary zone between active and passive, where passive execution of the safety

Page 150: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-134

function is accomplished through passive means (i.e., natural circulation), but the process is

initiated by active components (i.e., condensate valves actuation).

2. Scenario identification

2.A. Normal reactor operation:

During normal reactor operation the ICS is in standby mode with the steam line isolation valves

in fully open position and the drain valves in closed position. The steam line isolation valves are

open so that the HX tube bundles are at reactor system pressure. This allows condensate to build-

up in the HX up to piping high point and filling the condensate drain line, which are maintained at

a sub-cooled temperature by IC pool water. Live reactor steam is present in the steam supply line

up through the horizontal distribution branch piping that feeds steam to the IC HX upper headers,

[SBWR standard safety analysis report (1992)]. The vent and water makeup valves are also closed

during normal operation, unless during maintenance and water replenishment/pool cleanup.

2.B. Transient scenario:

The ICS is a standby high-pressure system that removes residual and decay heat from the reactor

pressure vessel (RPV) in the event of a reactor SCRAM in which, the reactor becomes isolated

from the main condenser, or if any other high-pressure abnormal condition exists. The ICS aids in

reactor vessel depressurization in the event that either the feedwater coolant injection or high-

pressure coolant injection system fails. In other words, the ICS passively removes core decay heat

following a reactor SCRAM from 100% full power operation and when the normal heat removal

system is unavailable [29]. Typically, the ICS is required to remove up to 4% (approx.) of rated

power which results from decay heat. Overall, ICS could be thought of as a pressure regulating

system during abnormal operation, and core decay heat removal system.

The ICS is initiated by any of the following signals/action:

1. RPV pressure above a specified threshold value;

2. Main steam line isolation valve (MSIV) closure fraction;

3. RPV water level below a specified threshold value;

4. Remote manual initiation by operators.

For instance, the ICS is automatically initiated if a high reactor pressure condition is sustained for

15 seconds (generic). The time delay prevents unnecessary system initiation during turbine trips

Page 151: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-135

[29]. Additionally, in many reactors design the ICS is automatically initiated on a low RPV water

level to aid in reactor depressurization for small line breaks. These initiating signals places the ICS

into operation by opening the condensate return line isolation valve, which results in draining of

the accumulated condensate. Saturated steam flows from the RPV to the HX tubes via steam supply

line and are condensed in the HX tubes. The condensate returns to the downcomer region of the

RPV by gravity via condensate drain line. The start-up process is sufficiently fast to limit the

reactor pressure rise resulting from reactor isolation to well below the pressure set-point of the

SRVs for all non-accident transient isolation event [SBWR design description]. The natural

circulation i.e., the buoyant force generated from coolant density difference dictates the coolant

flowrate in the IC loop. In other words, pressure in the condensate return line region determines

the coolant flowrate into the RPV. The opening of the condensate drain valve causes the liquid

level in the loop to drop and to increase the available tube surface area for vapor condensation

[Khan et al. (1992)]. This results in transfer of heat from the IC loop to the IC water pool, and the

heat transfer rate is dependent on the IC pool conditions, heat transfer characteristics of the IC

tubes, presence of non-condensable gases, state of the HX, etc. The extended operation of the ICS

i.e., a mission time of 72 hours is accomplished by replenishing water into the IC pool through

dedicated water makeup systems, thereby keeping the HX submerged in the IC water pool.

3. System mission

The primary objective/mission of the ICS includes:

1. Reactor pressure vessel depressurization;

2. Remove sensible and decay heat from the reactor for 72 hours (mission time);

3. Maintain fuel peak cladding temperature within design limits.

The above-mentioned objectives are typically achieved by employing a number of totally

independent IC loops for the purpose of redundancy. First, in trying to achieve the above mission,

the ICS also prevent unnecessary activation of safety relief valves (SRVs) by maintaining system

pressure below SRVs set-point. This in turn eliminate or mitigate loss-of-coolant accident (LOCA)

that may result from SRVs being fail-open. In other words, the ICS reduces the cycling frequency

(opening and closing) of the SRVs, and hence can decrease the probability of SRVs failure on

demand. Second, system depressurization under defined transient events can be achieved without

the loss of primary coolant inventory. Thus, the ICS remove excess sensible and core decay heat

Page 152: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-136

from the reactor in a passive way with minimal loss of coolant inventory when normal heat removal

systems are unavailable. Third, operation of high pressure coolant injection system could

potentially be eliminated or delay while maintaining the system pressure. Furthermore, in the event

of a small LOCA, the ICS is demanded to depressurize the RCS so that high pressure injection

system could be activated for coolant injection. The above three functions can be considered as

secondary objectives.

4. System success/failure criteria

The ICS failure criteria includes:

1. Failure to maintain the reactor coolant system pressure below threshold value;

2. Failure to remove specified decay and residual heat produced from the reactor, i.e., decay

heat removed < decay heat generated;

3. Failure to maintain peak fuel cladding temperature below threshold.

It may be observed that the above criteria are interrelated. For instance, criteria 2 can cause both

criteria 1 and 3 to occur, i.e., a failure to remove decay heat from the reactor will result to an

increase in system pressure as well as a heat up of the fuel which can eventually lead to fuel failure.

Also, a deviation in differential pressure can influence the condensate flowrate in the IC loop,

which can in turn affect the amount of heat transfer rate to the IC pool. These inter-dependencies

must be accounted for when developing the physical model of the system. For instance, the

success/failure criteria may be given in terms of performance indicator (PI) as:

𝑃𝐼 =Heat removed

Decay heat generated=

HR

DHG

(5-1)

𝑃𝐼 =

1; Ideal success state (HR = DHG)≥ 1; Not of interest for analysis

< 1; System failure or partial failure

(5-2)

A typical ICS is designed to remove four (4) percent of reactor rated power, which means that five

minutes after a scram and initiation of the ICS, the heat removal capacity of the IC system must

equal the decay heat production rate of the shutdown reactor. This will ensure that the system

pressure is below threshold limit. And note that, the reactor coolant system pressure is considered

as the state variable of interest for analysis purpose.

Page 153: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-137

5. The ICS components

The ICS consist of the following components: [Note that the ICS description is generic in nature

and does not imply any specific design]

a. Steam supply line piping (vapor phase)

The steam supply line connects the RPV to the HX tubes, i.e., live reactor steam enters the HX

tubes via steam line. This region remains under reactor coolant system pressure, and in vapor state.

The steam supply line is normally open with the steam line isolation valve in fully open position

during normal operation. The isolation valve is closed in case of leakage/rupture in the ICS piping

or rupture in the HX tubes. This region is typically guardpiped so that a break in the line is fully

contained.

b. The heat exchanger

The heat exchanger is a critical component of the ICS in which the live reactor steam is condensed

inside the HX tubes. The HX consist of a number of vertical placed tubes and is emerged in the

ICS water pool (shell side) creating an interface with the IC loop (tube side). The heat transfer

from one loop to the other takes place in the HX.

c. Condensate drain line (liquid phase)

The condensate return line begins from the lower header of the IC HX and are normally filled with

sub-cooled condensate. The condensate drain line consist of a series pair of isolation valves and a

parallel pair of condensate drain valves. The pressure in this region controls the performance of

the ICS during a transient event. The drain line piping ends at a dedicated condensate return nozzle

that are typically located at the mid-height of the RPV.

d. Condensate drain valves

The condensate drain valves are the most critical components of the ICS. The initiation, operation

and performance of the ICS is highly dependent on drain valves state. The coolant flowrate in the

IC loop and hence the heat transferred is controlled by the drain valve position. The drain valves

consist of the main and bypass valve that are connected in parallel. Typically, the main valve is a

motor operated with fail-as-is, and the bypass valve- a nitrogen piston operated with fail-open.

Page 154: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-138

e. Vent lines

The vent lines are typically installed at the lower and upper headers of the IC HXs with the main

purpose to purge air and non-condensable gases from the HXs which could significantly affect and

degrade the natural circulation of the coolant in the IC loop. They consist of a number of vent

valves (spring and motor operated) that are normally closed. The vent lines for each of the ICS

HXs are routed into the containment, and then to the suppression pool (typical advanced BWR

reactor designs). These lines are provided with two main and two bypass valves located in a series

of valves that are required to open at high reactor pressure during ICS operation.

f. Vent valves

During normal plant operation, air and non-condensable gases may accumulate in the ICS

condenser due to hydrogen buildup from water chemistry control additions and air entrained in the

feedwater. This could degrade the long-term heat removal capacity of the ICS. The purpose of the

vent valves is to remove accumulated air and non-condensable gases from the ICS loop during its

operation. The vent valves are installed in the vent line with a parallel configuration for redundancy

diversity. The vent valves operation is controlled by automatic logics as well as by operator manual

actuation. Venting is initiated whenever a combination of two signals is present:

• High RPV pressure;

• Operation/opening of the condensate drain valves.

The operating pressure of the venting unit is established below the set-point at which the lowest-

set SRVs will actuate. This help ensures that the SRVs will not actuate during a reactor isolation

transient even without operator intervention [29].

g. Water makeup systems

The purpose of the water makeup system is to supply water to the IC water pool during extended

ICS operation, and to ensure that the HX tubes remain covered at all time. A reduced water level

will lead to uncover of HX tubes and thus degrading the heat transfer rate from the IC loop to the

IC pool. Typically, a dedicated condensate water makeup system is in place, with the firewater

system as an alternate water supply system (typical advanced BWR designs).

Page 155: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-139

The ICS system characterization

The ICS hardware components is characterized into several units for stochastic modelling of the

system. The units are as follow:

(b) Condensate Drain Unit (DU)

(c) Heat Exchanger Unit (HXU)

(d) Vent Unit (VU)

(e) Water Make-up Unit (MWU)

(f) Primary boundary envelope (EF)

Of course, the ICS operation is significantly affected by the hardware unit states. However, there

are several physical parameters that significantly influence the ICS performance such as presence

of non-condensable gases, oxidation, pipe fouling and thermal stratification. Due to the

unavailability of failure data and their nature of influence on the system failure, these parameters

cannot be included in the stochastic system model at this point. Hence, only the hardware system

components are modelled using the Markov model. The author has proposed two approaches to

modelling the overall ICS system reliability:

1. Small Markov model and large fault tree: The individual system units are modelled using

Markov model whereas the overall system reliability is modelled using FT technique. It must

be underlined that the individual unit reliability is given in terms of probability of unit failure,

and hence quantification can be performed using any standard FT codes. The system reliability

is given in terms of probability of failure.

2. Large Markov model and small fault tree: In this approach, the individual units are modelled

in FT using the failure rate evaluation method. Note that the unit reliability is in terms of failure

rates, rather than the failure probability as in the first approach. Once the failure rates of the

individual units are obtained, the overall system reliability is modelled using time-dependent

Markov model. The system reliability is given in terms of probability of failure. This approach

is implemented in this thesis, since it allows for coupling the model with CCMT.

The key difference between the two approaches is that the first approach uses probability

evaluation method to determine the individual unit’s reliability, whereas the second approach uses

failure rate evaluation for individual units. However, both methods provide the result in terms of

Page 156: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-140

probability of failure. The second approach have an advantage over the first since it can yield a

time-dependent model of the overall system, and the interactions among the units. In a way, these

two approaches can be used as a complementary to each other and add credit to the predicted

system reliability. The two (2) approaches are used to model the individual units of the ICS. The

results obtained from the individual unit analysis is utilized for IC system reliability assessment.

The computation of the time-dependent failure rate under a logic AND gate is determined as:

𝜆𝑆(𝑡) =∑ 𝜆𝑖(𝛼𝑖 − 1)𝑛𝑖=1

𝛼𝑖𝑛𝑖=1 − 1

(5-3)

Where; 𝛼𝑖 =1

(1−𝑒−(𝜆𝑖𝑡))

𝜆𝑖 = individual component failure rate

𝑛 = number of components connected in parallel and are within the unit

𝜆𝑆(𝑡) = time-dependent unit failure rate

For two identical components in parallel with the same failure rate given by 𝜆, the time dependent

system failure rate is:

𝜆𝑆(𝑡) =2𝜆

𝛼 + 1

(5-4)

The above relations are implemented throughout this sub-section to determine the time-dependent

unit failure rate, besides the Markov model that directly yields the unit failure probability.

A. Condensate Drain Unit

The condensate drain unit consist of two drain valves i.e., the main drain valve and bypass valve

connected in parallel for redundancy and diversity of the unit. Each of the valve is considered to

have three (3) possible states as shown in Table 43. It is assumed that the main and bypass valve

have the same modes of failure including common cause failure (CCF) and failure rate (for

simplicity).

Page 157: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-141

Table 43: Failure modes and rate of the main/bypass valve

Valves Failure modes Failure rate (per yr)

Main/bypass

drain valve

Normal −

Failed to remain open (𝜆2𝐷) 8.6 × 10−09

Failed-to-open (𝜆1𝐷) 1.4 × 10−06

Failed-to-open due to CCF (𝜆1,𝐶𝐶𝐹𝐷 ) 1.2 × 10−04

Failed to remain open due to CCF (𝜆2,𝐶𝐶𝐹𝐷 ) 7.2 × 10−07

Enumerating all possible component state combination (See Table 44), the drain unit can have nine

(9) possible states (See Table 45). Furthermore, state merging can be performed to reduce the

overall unit states. For the drain unit, the nine (9) states are merged into two (2) unit states, i.e.,

normal or failure. The unit is considered failed if both the units are in any failed or partially failed

state, and a unit normal/operational state if any of the valve are in normal state.

Table 44: Possible drain unit state combination

Unit state combination Unit States

(𝑛) Main valve states Bypass valve states

Normal Normal Normal

Normal Fail-to-open Normal

Normal Failed to remain open Normal

Fail-to-open Normal Normal

Fail-to-open Fail-to-open Failure

Fail-to-open Failed to remain open Failure

Failed to remain open Normal Normal

Failed to remain open Fail-to-open Failure

Failed to remain open Failed to remain open Failure

Page 158: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-142

Table 45: Unit state ordering and the inclusion of CCF

Unit State Combination CCF of the

valves Unit States

(𝑛) Main valve states Bypass valve states

𝑛1 = 0 𝑛2 = 0 - 1

𝑛1 = 0 𝑛2 = 1 - 2

𝑛1 = 0 𝑛2 = 2 - 3

𝑛1 = 1 𝑛2 = 0 - 4

𝑛1 = 1 𝑛2 = 1 CCF 5

𝑛1 = 1 𝑛2 = 2 - 6

𝑛1 = 2 𝑛2 = 0 - 7

𝑛1 = 2 𝑛2 = 1 - 8

𝑛1 = 2 𝑛2 = 2 CCF 9

Note: 0 = component normal; 1 = component fail-to-open; 2 = component fail to remain open.

The Markov state transition diagram of the condensate drain unit is given in Figure 29.

Figure 29: Markov transition diagram of the condensate drain unit

The ODEs from the above transition diagram are:

𝑑𝑃1(𝑡)

𝑑𝑡= −(𝜆1

𝐵 + 𝜆2𝐵 + 𝜆1

𝐷 + 𝜆2𝐷 + 𝜆𝐶𝐶𝐹

𝑓𝑡𝑜+ 𝜆𝐶𝐶𝐹

𝐼𝑉𝐶 )𝑃1(𝑡)

𝑑𝑃2(𝑡)

𝑑𝑡= 𝜆1

𝐵𝑃1(𝑡) − (𝜆1𝐷 + 𝜆2

𝐷 + 𝜆𝐶𝐶𝐹𝑓𝑡𝑜

)𝑃2(𝑡)

Page 159: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-143

𝑑𝑃3(𝑡)

𝑑𝑡= 𝜆2

𝐵𝑃1(𝑡) − (𝜆1𝐷 + 𝜆2

𝐷 + 𝜆𝐶𝐶𝐹𝐼𝑉𝐶 )𝑃3(𝑡)

𝑑𝑃4(𝑡)

𝑑𝑡= 𝜆1

𝐷𝑃1(𝑡) − (𝜆1𝐵 + 𝜆2

𝐵 + 𝜆𝐶𝐶𝐹𝑓𝑡𝑜

)𝑃4(𝑡)

𝑑𝑃7(𝑡)

𝑑𝑡= 𝜆2

𝐷𝑃1(𝑡) − (𝜆1𝐵 + 𝜆2

𝐵 + 𝜆𝐶𝐶𝐹𝐼𝑉𝐶 )𝑃7(𝑡)

𝑑𝑃5(𝑡)

𝑑𝑡= 𝜆𝐶𝐶𝐹

𝑓𝑡𝑜𝑃1(𝑡) + (𝜆1

𝐷 + 𝜆𝐶𝐶𝐹𝑓𝑡𝑜

)𝑃2(𝑡) + (𝜆1𝐵 + 𝜆𝐶𝐶𝐹

𝑓𝑡𝑜)𝑃4(𝑡)

𝑑𝑃8(𝑡)

𝑑𝑡= 𝜆2

𝐷𝑃2(𝑡) + 𝜆1𝐵𝑃7(𝑡)

𝑑𝑃6(𝑡)

𝑑𝑡= 𝜆1

𝐷𝑃3(𝑡) + 𝜆2𝐵𝑃4(𝑡)

𝑑𝑃9(𝑡)

𝑑𝑡= 𝜆𝐶𝐶𝐹

𝐼𝑉𝐶 𝑃1(𝑡) + (𝜆2𝐵 + 𝜆𝐶𝐶𝐹

𝐼𝑉𝐶 )𝑃7(𝑡) + (𝜆2𝐷 + 𝜆𝐶𝐶𝐹

𝐼𝑉𝐶 )𝑃3(𝑡)

(5-5)

The solution of the abo2ve ODEs is obtained using finite difference method implemented in

Fortran95 code. Sample unit state transition probabilities for 4 time steps are shown in Table 46.

Table 46: Condensate drain unit state transition probabilities

Unit states 𝑘 = 0 𝑘 = 1 𝑘 = 2 𝑘 = 3

𝑃1(𝑘∆𝑡) 1.00 0.998 0.997 0.996

𝑃2(𝑘∆𝑡) 0 1.4E-05 2.797E-05 4.19E-05

𝑃3(𝑘∆𝑡) 0 8.6E-08 1.72E-07 2.58E-07

𝑃4(𝑘∆𝑡) 0 1.4E-05 2.79E-05 4.19E-05

𝑃5(𝑘∆𝑡) 0 1.2E-03 2.39E-03 3.59E-03

𝑃6(𝑘∆𝑡) 0 0 1.04E-10 3.13E-10

𝑃7(𝑘∆𝑡) 0 8.6E-08 1.72E-07 2.58E-07

𝑃8(𝑘∆𝑡) 0 0 2.41E-12 7.22E-12

𝑃9(𝑘∆𝑡) 0 7.2E-06 1.44E-05 2.16E-05

Sum 1.00 1.00 1.00 1.00

A state merging on Table 46 gives Table 47.

2 Note that the ODEs for individual units are grouped and depicted by a single equation for ease of representation

Page 160: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-144

Table 47: Merged condensate drain unit state

Unit state 𝑘 = 0 𝑘 = 1 𝑘 = 2 𝑘 = 3

Unit normal 1.00 0.998 0.997 0.996

Unit failure 0 1.21E-03 2.41E-03 3.62E-03

The FT of the condensate drain unit is depicted in Figure 30.

Since the failure rates of main and bypass valve are identical, the unit failure rate is given as:

𝜆𝑆(𝑡) =2𝜆

𝛼 + 1

(5-6)

Where; 𝛼 =1

(1−𝑒−(𝜆𝑡))

The top event or failure rate of the condensate drain unit is, 𝜆𝐷_𝑈𝑛𝑖𝑡(𝑡) = 1.21 × 10−04𝑝𝑒𝑟𝑦𝑟.

Figure 30: Fault tree of the condensate drain unit

B. Vent Unit

The vent unit consist of two vent valves connected in parallel for redundancy of the unit. Each of

the valve (component level) is considered to have three (3) possible states (See Table 48). A

common cause failure (CCF) of the valves are considered, and the vent valves are considered to

have the same failure rate (for simplicity).

Page 161: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-145

Table 48: Failure modes and rate of the vent valve

Component Failure modes Failure rate (per yr)

Vent valve Normal −

Failed closed (𝜆𝑉𝑓𝑐) 1.7 × 10−08

Failed open (𝜆𝑉𝑓𝑜) 2.9 × 10−06

Failed closed due to CCF (𝜆𝐶𝐶𝐹𝑓𝑐

) 7.2 × 10−07

Failed open due to CCF (𝜆𝐶𝐶𝐹𝑓𝑜

) 1.2 × 10−04

The overall vent unit is characterized and merged into three (3) possible states: (1) Normal; (2)

Partial unit failure; and (3) Total unit failure. The set of possible states of the venting unit or

component state combination is found by enumerating all the possible modes of failure of the two

vent valves (See Table 49).

Table 49: Unit state numbering and the inclusion of CCF

Component State Combination Unit States

(𝑛) Valve-1 state Valve-2 state

𝑛1 = 0 𝑛2 = 0 1

𝑛1 = 0 𝑛2 = 1 2

𝑛1 = 0 𝑛2 = 2 3

𝑛1 = 1 𝑛2 = 0 4

𝑛1 = 1 𝑛2 = 1 5

𝑛1 = 1 𝑛2 = 2 6

𝑛1 = 2 𝑛2 = 0 7

𝑛1 = 2 𝑛2 = 1 8

𝑛1 = 2 𝑛2 = 2 9

A partial unit failure is considered for the venting unit due to the ICS performance sensitivity to

the presence of non-condensable gas fraction. Partial failure implies the vent valves getting fail in

any intermediate position (e.g., 50%, 60%, 40%, etc.), which provide a flexibility to observe the

system response under varying system configuration. Of course, both the valves failing open or

Page 162: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-146

close will result in a unit failure. The Markov transition diagram for the vent unit is given in Figure

31. Note that, V= vent valve 1, and B= vent valve 2.

Figure 31: Markov transition diagram of the vent unit

The ODEs obtained from the state transition diagram are:

𝑑𝑃1(𝑡)

𝑑𝑡= −(𝜆1

𝐵 + 𝜆2𝐵 + 𝜆1

𝑉 + 𝜆2𝑉 + 𝜆𝐶𝐶𝐹

𝑓𝑐+ 𝜆𝐶𝐶𝐹

𝑓𝑜)𝑃1(𝑡)

𝑑𝑃2(𝑡)

𝑑𝑡= 𝜆1

𝐵𝑃1(𝑡) − (𝜆1𝑉 + 𝜆2

𝑉 + 𝜆𝐶𝐶𝐹𝑓𝑐

)𝑃2(𝑡)

𝑑𝑃3(𝑡)

𝑑𝑡= 𝜆2

𝐵𝑃1(𝑡) − (𝜆1𝑉 + 𝜆2

𝑉 + 𝜆𝐶𝐶𝐹𝑓𝑜

)𝑃3(𝑡)

𝑑𝑃4(𝑡)

𝑑𝑡= 𝜆1

𝑉𝑃1(𝑡) − (𝜆1𝐵 + 𝜆2

𝐵 + 𝜆𝐶𝐶𝐹𝑓𝑐

)𝑃4(𝑡)

𝑑𝑃7(𝑡)

𝑑𝑡= 𝜆2

𝑉𝑃1(𝑡) − (𝜆1𝐵 + 𝜆2

𝐵 + 𝜆𝐶𝐶𝐹𝑓𝑜

)𝑃7(𝑡)

𝑑𝑃5(𝑡)

𝑑𝑡= 𝜆𝐶𝐶𝐹

𝑓𝑐𝑃1(𝑡) + (𝜆1

𝑉 + 𝜆𝐶𝐶𝐹𝑓𝑐

)𝑃2(𝑡) + (𝜆1𝐵 + 𝜆𝐶𝐶𝐹

𝑓𝑐)𝑃4(𝑡)

𝑑𝑃8(𝑡)

𝑑𝑡= 𝜆2

𝑉𝑃2(𝑡) + 𝜆1𝐵𝑃7(𝑡)

𝑑𝑃6(𝑡)

𝑑𝑡= 𝜆1

𝑉𝑃3(𝑡) + 𝜆2𝐵𝑃4(𝑡)

𝑑𝑃9(𝑡)

𝑑𝑡= 𝜆𝐶𝐶𝐹

𝑓𝑜𝑃1(𝑡) + (𝜆2

𝐵 + 𝜆𝐶𝐶𝐹𝑓𝑜

)𝑃7(𝑡) + (𝜆2𝑉 + 𝜆𝐶𝐶𝐹

𝑓𝑜)𝑃3(𝑡)

(5-7)

Performing state merging on the solution of the above ODEs with normal unit state at 𝑡 = 0 yields

Table 50. Note only sample four (4) time steps is shown for illustration purpose.

Page 163: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-147

Table 50: Vent unit states transition probabilities

Unit state 𝑘 = 0 𝑘 = 1 𝑘 = 2 𝑘 = 3

Normal 1.00 0.998 0.997 0.996

Partial failure 0 0.00 1.97E-11 5.91E-11

Total failure 0 1.21E-03 2.413E-03 3.62E-03

The FT for the vent unit using failure rate evaluation method is performed for two (2) modes, i.e.,

the unit total failure and partial failure. The total unit failure rate obtained from the FT is:

Unit total failure = (𝜆𝑉𝑓𝑐𝜆𝐵𝑓𝑐+ 𝜆𝐶𝐶𝐹

𝑓𝑐) + (𝜆𝑉

𝑓𝑜𝜆𝐵𝑓𝑜+ 𝜆𝐶𝐶𝐹

𝑓𝑜) (5-8)

= 1.21 × 10−04𝑝𝑒𝑟𝑦𝑟

Whereas, the unit partial failure rate is:

Unit partial failure= (𝜆𝑉𝑓𝑐𝜆𝐵𝑓𝑜) + (𝜆𝑉

𝑓𝑜𝜆𝐵𝑓𝑐) (5-9)

Unit partial failure = 1.62 × 10−15𝑝𝑒𝑟𝑦𝑟

C. HX Unit

The HX unit consist of a single U-tube heat exchanger immersed in a pool of water. The HX is

considered to have the failure modes and rates as shown in Table 51.

Table 51: Failure modes and rate of the heat exchanger

Component Failure modes Failure rate (per yr)

Heat

exchanger

Normal −

Single pipe rupture 2.63 × 10−06

Multiple pipe rupture 2.63 × 10−07

Single pipe plugging 2.63 × 10−06

Multiple pipe plugging 2.63 × 10−07

The HX creates the interface between the two loops, and is the most critical component of the ICS,

in that, heat transfer takes place from the reactor to the IC pool. A change in the HX configuration

could significantly affect the ICS performance, and hence the system pressure. Investigation was

performed into the possible number of states that the HX can take. Nine (9) possible states for the

Page 164: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-148

HX is considered, and a qualitative outcome of the state combinations and individual failure modes

was made (See Table 52).

Table 52: Possible states of the HX unit

HX Unit state combination Unit states

(𝑛)

System states

Normal (𝑛1 = 0) 1 Normal

Single tube rupture (𝑛1 = 1) 2 Partial failure

Multiple tube rupture (𝑛1 = 2) 3 System failure

Single pipe plugging (𝑛1 = 3) 4 Partial failure

Multiple pipe plugging (𝑛1 = 4) 5 System failure

Single tube rupture (𝑛1 = 1) and Single

pipe plugging (𝑛1 = 3)

6 Partial failure

Single tube rupture (𝑛1 = 1) and

Multiple pipe plugging (𝑛1 = 4)

7 System failure

Multiple tube rupture (𝑛1 = 2) and

Single pipe plugging (𝑛1 = 3)

8 System failure

Multiple tube rupture (𝑛1 = 2) and

Multiple pipe plugging (𝑛1 = 4)

9 System failure

This approach enables one to reduce the number of states, and thus to avoid state space explosion

phenomenon. For instance, a single tube rupture or plugging may not necessary lead to a total

system failure. Hence, a partial system failure is considered, and the individual failure modes are

grouped into one state assuming that the consequences of both the failure modes are similar.

Merging all possible states based on their consequences, the HX unit is characterized into three (3)

possible states: normal, partial failure and total unit failure.

Some of the above unit states are highly unlikely but have non-zero probabilities. For instance,

multiple tube rupture and multiple pipe plugging simultaneously occurring is highly unlikely but

have occurred in NPPs during the past decade. Note that an independence is considered between

tube plugging and rupture, which is realistic. The Markov state transition diagram for the HX unit

is given in Figure 32. Even though, the HX unit consist only of one component, the state transition

Page 165: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-149

diagram is more complicated. This is due to the possible states as well as the possible state

evolution of the HX unit. For instance, a single tube rupture can evolve into a multiple tube rupture

state, or a single tube rupture along with single tube plugging can evolve into multiple tube rupture

with single tube plugging.

Figure 32: Markov state transition diagram for the HX unit

The resulting ODEs from the state transition diagram are:

𝑑𝑃1(𝑡)

𝑑𝑡= −(𝜆𝐻𝑋

𝑆𝑇𝑅 + 𝜆𝐻𝑋𝑀𝑇𝑅 + 𝜆𝐻𝑋

𝑆𝑃𝑃 + 𝜆𝐻𝑋𝑀𝑃𝑃)𝑃1(𝑡)

𝑑𝑃2(𝑡)

𝑑𝑡= 𝜆𝐻𝑋

𝑆𝑇𝑅𝑃1(𝑡) − (𝜆𝐻𝑋𝑆𝑇𝑅 + 𝜆𝐻𝑋

𝑆𝑃𝑃 + 𝜆𝐻𝑋𝑀𝑃𝑃)𝑃2(𝑡)

𝑑𝑃3(𝑡)

𝑑𝑡= 𝜆𝐻𝑋

𝑀𝑇𝑅𝑃1(𝑡) + 𝜆𝐻𝑋𝑆𝑇𝑅𝑃2(𝑡) − (𝜆𝐻𝑋

𝑆𝑃𝑃 + 𝜆𝐻𝑋𝑀𝑃𝑃)𝑃3(𝑡)

𝑑𝑃4(𝑡)

𝑑𝑡= 𝜆𝐻𝑋

𝑆𝑃𝑃𝑃1(𝑡) − (𝜆𝐻𝑋𝑆𝑇𝑅 + 𝜆𝐻𝑋

𝑀𝑇𝑅 + 𝜆𝐻𝑋𝑆𝑃𝑃)𝑃4(𝑡)

𝑑𝑃5(𝑡)

𝑑𝑡= 𝜆𝐻𝑋

𝑆𝑃𝑃𝑃2(𝑡) + 𝜆𝐻𝑋𝑆𝑇𝑅𝑃4(𝑡) − (𝜆𝐻𝑋

𝑆𝑇𝑅 + 𝜆𝐻𝑋𝑆𝑃𝑃)𝑃5(𝑡)

𝑑𝑃6(𝑡)

𝑑𝑡= 𝜆𝐻𝑋

𝑠𝑝𝑝𝑃3(𝑡) + 𝜆𝐻𝑋

𝑚𝑡𝑟𝑃4(𝑡) + 𝜆𝐻𝑋𝑠𝑡𝑟𝑃5(𝑡) − 𝜆𝐻𝑋

𝑠𝑝𝑝𝑃6(𝑡)

𝑑𝑃7(𝑡)

𝑑𝑡= 𝜆𝐻𝑋

𝑀𝑃𝑃𝑃1(𝑡) + 𝜆𝐻𝑋𝑆𝑃𝑃𝑃4(𝑡) − (𝜆𝐻𝑋

𝑆𝑇𝑅 + 𝜆𝐻𝑋𝑀𝑇𝑅)𝑃7(𝑡)

𝑑𝑃8(𝑡)

𝑑𝑡= 𝜆𝐻𝑋

𝑚𝑝𝑝𝑃2(𝑡) + 𝜆𝐻𝑋

𝑠𝑝𝑝𝑃5(𝑡) + 𝜆𝐻𝑋

𝑠𝑡𝑟𝑃7(𝑡) − 𝜆𝐻𝑋𝑠𝑡𝑟𝑃8(𝑡)

Page 166: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-150

𝑑𝑃9(𝑡)

𝑑𝑡= 𝜆𝐻𝑋

𝑚𝑝𝑝𝑃3(𝑡) + 𝜆𝐻𝑋

𝑠𝑝𝑝𝑃6(𝑡) + 𝜆𝐻𝑋

𝑚𝑡𝑟 𝑃7(𝑡) + 𝜆𝐻𝑋𝑠𝑡𝑟𝑃8(𝑡)

(5-10)

Performing state merging on the solution of the ODEs for four (4) time steps yields Table 53.

Table 53: The overall HX unit state probabilities

Unit state 𝑘 = 0 𝑘 = 1 𝑘 = 2 𝑘 = 3

Normal 1.00 0.9999 0.9998 0.9998

Partial failure 0 2.89E-05 5.78E-05 8.67E-05

Total failure 0 5.26E-06 1.05E-05 1.58E-05

A failure rate evaluation for partial failure and total failure of the HX is performed to determine

the time dependent failure rate that is required to construct and compute the overall ICS state

probabilities. Unit state 2, 4 and 6 in Table 52 implies a partial failure of the unit, hence the failure

rate of the unit to be in partially failed state can simply be determined by.

Partial unit failure= {(𝑛1 = 1) + (𝑛1 = 3) + {(𝑛1 = 1)𝐴𝑁𝐷(𝑛1 = 3)}} (5-11)

= 5.26 × 10−06𝑝𝑒𝑟𝑦𝑟

Unit states 3, 5, 7, 8 and 9 implies a total unit failure. Hence, a total unit failure probability is:

Total unit failure= {(𝑛1 = 2) + (𝑛1 = 4) + {(𝑛1 = 1)𝐴𝑁𝐷(𝑛1 = 4)} + {(𝑛1 = 2)𝐴𝑁𝐷(𝑛1 = 3)} +

{(𝑛1 = 2)𝐴𝑁𝐷(𝑛1 = 4)}} (5-12)

= 5.26 × 10−07𝑝𝑒𝑟𝑦𝑟

D. Water Make-up Unit

The water makeup-up unit consist of two redundant and diverse systems as follow:

• Main water make-up system (e.g., condensate)

• Alternate water make-up system (e.g., Firewater)

The two systems are treated as two components connected in parallel with each component have

three possible states: (1) Normal; (2) Fail-to-open; and (3) Fail-to-remain-open (See Table 54). No

common cause failure (CCF) of the system are considered due to the nature of diversity and

independence between the two systems. The failure modes and their respective failure rate are:

Page 167: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-151

Table 54: Failure modes and rate of the vent valve

Component Failure modes Failure rate (per yr)

Main system Fail to open (𝜆1𝑀) 1.2 × 10−03

Fail to remain open (𝜆2𝑀) 1.0 × 10−04

Alternate system Fail to open (𝜆1𝐴) 1.4 × 10−03

Fail to remain open (𝜆2𝐴) 1.0 × 10−04

The overall water make-up unit is characterized and merged into two (2) possible system states:

(1) unit normal; and (2) unit failure. The set of possible water make-up unit states is found by

enumerating all possible modes of failure of the main and alternate makeup system (See Table 55).

Table 55: Unit state combinations and ordering (notations)

Unit State Combination Unit States

(𝑛) Main valve states Alternate valve states

𝑛1 = 0 𝑛2 = 0 1

𝑛1 = 0 𝑛2 = 1 2

𝑛1 = 0 𝑛2 = 2 3

𝑛1 = 1 𝑛2 = 0 4

𝑛1 = 1 𝑛2 = 1 5

𝑛1 = 1 𝑛2 = 2 6

𝑛1 = 2 𝑛2 = 0 7

𝑛1 = 2 𝑛2 = 1 8

𝑛1 = 2 𝑛2 = 2 9

The Markov state transition diagram for the water makeup unit is depicted in Figure 33.

Page 168: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-152

Figure 33: Markov state transition diagram of the water makeup unit

The resulting ODEs from the above state transition diagram are:

𝑑𝑃1(𝑡)

𝑑𝑡= −(𝜆1

𝐴 + 𝜆2𝐴 + 𝜆1

𝑀 + 𝜆2𝑀)𝑃1(𝑡)

𝑑𝑃2(𝑡)

𝑑𝑡= 𝜆1

𝐴𝑃1(𝑡) − (𝜆1𝑀 + 𝜆2

𝑀)𝑃2(𝑡)

𝑑𝑃3(𝑡)

𝑑𝑡= 𝜆2

𝐴𝑃1(𝑡) − (𝜆1𝑀 + 𝜆2

𝑀)𝑃3(𝑡)

𝑑𝑃4(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑃1(𝑡) − (𝜆1𝐴 + 𝜆2

𝐴)𝑃4(𝑡)

𝑑𝑃5(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑃2(𝑡) + 𝜆1𝐴𝑃4(𝑡)

𝑑𝑃6(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑃3(𝑡) + 𝜆2𝐴𝑃4(𝑡)

𝑑𝑃7(𝑡)

𝑑𝑡= 𝜆2

𝑀𝑃1(𝑡) − (𝜆1𝐴 + 𝜆2

𝐴)𝑃7(𝑡)

𝑑𝑃8(𝑡)

𝑑𝑡= 𝜆2

𝑀𝑃2(𝑡) + 𝜆1𝐴𝑃7(𝑡)

𝑑𝑃9(𝑡)

𝑑𝑡= 𝜆2

𝑀𝑃3(𝑡) + 𝜆2𝐴𝑃7(𝑡)

(5-13)

Analytical solution of the above ODEs is presented below. This was done to check the steady state

behavior of the unit states.

𝑃2(𝑡) =𝜆1𝐴

𝜆1𝐴+𝜆2

𝐴 𝑒−(𝜆1

𝑀+𝜆2𝑀)𝑡 −

𝜆1𝐴

𝜆1𝐴+𝜆2

𝐴 𝑒−(𝜆1

𝐴+𝜆2𝐴+𝜆1

𝑀+𝜆2𝑀)𝑡 (5-14)

𝑃5(𝑡) =𝜆1𝐴𝜆1

𝑀

𝜆1𝐴𝜆1

𝑀 + 𝜆1𝐴𝜆2

𝑀 + 𝜆2𝐴𝜆1

𝑀 + 𝜆2𝐴𝜆2

𝑀 +𝜆1𝐴𝜆1

𝑀

𝜆1𝐴𝜆1

𝑀 + 𝜆1𝐴𝜆2

𝑀 + 𝜆2𝐴𝜆1

𝑀 + 𝜆2𝐴𝜆2

𝑀 𝑒−(𝜆1𝐴+𝜆2

𝐴+𝜆1𝑀+𝜆2

𝑀)𝑡

−𝜆1𝐴𝜆1

𝑀

𝜆1𝐴𝜆1

𝑀+𝜆1𝐴𝜆2

𝑀+𝜆2𝐴𝜆1

𝑀+𝜆2𝐴𝜆2

𝑀 𝑒−(𝜆1

𝐴+𝜆2𝐴)𝑡 −

𝜆1𝐴𝜆1

𝑀

𝜆1𝐴𝜆1

𝑀+𝜆1𝐴𝜆2

𝑀+𝜆2𝐴𝜆1

𝑀+𝜆2𝐴𝜆2

𝑀 𝑒−(𝜆1

𝑀+𝜆2𝑀)𝑡 (5-15)

Page 169: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-153

𝑃9(𝑡) =𝜆2𝐴𝜆2

𝑀

𝜆1𝐴𝜆1

𝑀 + 𝜆1𝐴𝜆2

𝑀 + 𝜆2𝐴𝜆1

𝑀 + 𝜆2𝐴𝜆2

𝑀 +𝜆2𝐴𝜆2

𝑀

𝜆1𝐴𝜆1

𝑀 + 𝜆1𝐴𝜆2

𝑀 + 𝜆2𝐴𝜆1

𝑀 + 𝜆2𝐴𝜆2

𝑀 𝑒−(𝜆1𝐴+𝜆2

𝐴+𝜆1𝑀+𝜆2

𝑀)𝑡

−𝜆2𝐴𝜆2

𝑀

𝜆1𝐴𝜆1

𝑀+𝜆1𝐴𝜆2

𝑀+𝜆2𝐴𝜆1

𝑀+𝜆2𝐴𝜆2

𝑀 𝑒−(𝜆1

𝐴+𝜆2𝐴)𝑡 −

𝜆2𝐴𝜆2

𝑀

𝜆1𝐴𝜆1

𝑀+𝜆1𝐴𝜆2

𝑀+𝜆2𝐴𝜆1

𝑀+𝜆2𝐴𝜆2

𝑀 𝑒−(𝜆1

𝑀+𝜆2𝑀)𝑡 (5-16)

For analyzing the steady state behavior, take the case for 𝑃2(𝑡) as 𝑡 → ∞.

𝑃2(𝑡) =𝜆1𝐴

𝜆1𝐴+𝜆2

𝐴 𝑒−∞ −

𝜆1𝐴

𝜆1𝐴+𝜆2

𝐴 𝑒−∞ (5-17)

𝑃2(𝑡) = 0

Again, take the case for 𝑃5(𝑡) as 𝑡 → ∞.

𝑃5(𝑡) =𝜆1𝐴𝜆1

𝑀

𝜆1𝐴𝜆1

𝑀+𝜆1𝐴𝜆2

𝑀+𝜆2𝐴𝜆1

𝑀+𝜆2𝐴𝜆2

𝑀 (5-18)

𝑃5(𝑡) = 0.862 (steady state value)

A state merging on the solution of the ODEs yields two unit state transition probabilities. Sample

state probabilities for four (4) time steps are given in Table 56.

Table 56: Overall makeup water unit states

Unit state 𝑘 = 0 𝑘 = 1 𝑘 = 2 𝑘 = 3

Normal 1.00 1.0E+00 0.999 0.998

Unit failure 0 0 3.90E-04 1.15E-03

The failure rate evaluation from the FT of the water makeup unit can simply be given by:

Unit failure rate = 𝜆1𝑀𝜆1

𝐴 + 𝜆2𝑀𝜆1

𝐴 + 𝜆1𝑀𝜆2

𝐴 + 𝜆2𝑀𝜆2

𝐴 (5-19)

= 3.21 × 10−08𝑝𝑒𝑟𝑦𝑟

E. Envelope failure (EF) or pipe break/rupture

Envelope failure implies a degraded state of the primary IC loop. A failure of IC loop will result

in a loss of coolant inventory from the RCS. Envelope failure can be categorized based on the

break size and location, however for modelling simplicity, only two state are considered, i.e.,

normal or failure. Failure rate of the envelope is 𝜆𝐸𝐹 = 2.74 × 10−04𝑝𝑒𝑟𝑦𝑟. Therefore, unit states

probability via Markov model is:

𝑃0(𝑡) = 𝑒−𝜆𝐸𝐹.𝑡 and 𝑃1(𝑡) = (1 − 𝑒−𝜆𝐸𝐹.𝑡) (5-20)

Page 170: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-154

The probability of EF being in a failed or pipe break state can easily be found using the solution

for 𝑃1(𝑡) at any point in time. Note that the mission time considered for the analysis is 72 hours.

Markov Model of the Benchmark Passive ICS

This sub-section integrates the results to predict the overall ICS reliability. The units are modelled

as a single entity or a macro-component thus merging the system states. This significantly reduces

the computation time and complexity, without compromising the model and prediction accuracy.

The number of possible states that the ICS can take any point in space is first determined from the

number of unit possible states. From the previous sections, the individual units have the following

number of states:

1. Vent unit: 3 states

2. Makeup water unit: 2 states

3. Drain unit: 2 states

4. System envelope: 2 states

5. HX unit: 3 states

Therefore, the number of possible states that the ICS can take is 5 × 12 = 72discrete system states

(See Table 58). For ease of identification and analysis purpose, the individual unit’s notations and

numbering is categorized as shown in Table 57.

Table 57: Individual unit notation and numbering

Units Unit number

Condensate Drain Unit Unit-1

Vent Unit Unit-2

HX Unit Unit-3

Makeup Water Unit Unit-4

Envelope Failure Unit-5

Page 171: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-155

Table 58: Possible individual unit state combinations

Individual Unit State Combination System States

(𝑛) Drain unit

(𝑛1) Vent unit

(𝑛2) HX unit

(𝑛3) MK unit

(𝑛4) EF (𝑛5)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 0 𝑛4 = 0 𝑛5 = 0 0 𝑃0(𝑡)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 0 𝑛4 = 0 𝑛5 = 1 1 𝑃1(𝑡)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 0 𝑛4 = 1 𝑛5 = 0 2 𝑃2(𝑡)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 0 𝑛4 = 1 𝑛5 = 1 3 𝑃3(𝑡)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 1 𝑛4 = 0 𝑛5 = 0 4 𝑃4(𝑡)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 1 𝑛4 = 0 𝑛5 = 1 5 𝑃5(𝑡)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 1 𝑛4 = 1 𝑛5 = 0 6 𝑃6(𝑡)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 1 𝑛4 = 1 𝑛5 = 1 7 𝑃7(𝑡)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 2 𝑛4 = 0 𝑛5 = 0 8 𝑃8(𝑡)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 2 𝑛4 = 0 𝑛5 = 1 9 𝑃9(𝑡)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 2 𝑛4 = 1 𝑛5 = 0 10 𝑃10(𝑡)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 2 𝑛4 = 1 𝑛5 = 1 11 𝑃11(𝑡)

𝑛1 = 0 𝑛2 = 1 𝑛3 = 0 𝑛4 = 0 𝑛5 = 0 12 𝑃12(𝑡)

𝑛1 = 0 𝑛2 = 1 𝑛3 = 0 𝑛4 = 0 𝑛5 = 1 13 𝑃13(𝑡)

𝑛1 = 0 𝑛2 = 1 𝑛3 = 0 𝑛4 = 1 𝑛5 = 0 14 𝑃14(𝑡)

𝑛1 = 0 𝑛2 = 1 𝑛3 = 0 𝑛4 = 1 𝑛5 = 1 15 𝑃15(𝑡)

𝑛1 = 0 𝑛2 = 1 𝑛3 = 1 𝑛4 = 0 𝑛5 = 0 16 𝑃16(𝑡)

𝑛1 = 0 𝑛2 = 1 𝑛3 = 1 𝑛4 = 0 𝑛5 = 1 17 𝑃17(𝑡)

𝑛1 = 0 𝑛2 = 1 𝑛3 = 1 𝑛4 = 1 𝑛5 = 0 18 𝑃18(𝑡)

𝑛1 = 0 𝑛2 = 1 𝑛3 = 1 𝑛4 = 1 𝑛5 = 1 19 𝑃19(𝑡)

𝑛1 = 0 𝑛2 = 1 𝑛3 = 2 𝑛4 = 0 𝑛5 = 0 20 𝑃20(𝑡)

𝑛1 = 0 𝑛2 = 1 𝑛3 = 2 𝑛4 = 0 𝑛5 = 1 21 𝑃21(𝑡)

𝑛1 = 0 𝑛2 = 1 𝑛3 = 2 𝑛4 = 1 𝑛5 = 0 22 𝑃22(𝑡)

𝑛1 = 0 𝑛2 = 1 𝑛3 = 2 𝑛4 = 1 𝑛5 = 1 23 𝑃23(𝑡)

𝑛1 = 0 𝑛2 = 2 𝑛3 = 0 𝑛4 = 0 𝑛5 = 0 24 𝑃24(𝑡)

𝑛1 = 0 𝑛2 = 2 𝑛3 = 0 𝑛4 = 0 𝑛5 = 1 25 𝑃25(𝑡)

Page 172: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-156

𝑛1 = 0 𝑛2 = 2 𝑛3 = 0 𝑛4 = 1 𝑛5 = 0 26 𝑃26(𝑡)

𝑛1 = 0 𝑛2 = 2 𝑛3 = 0 𝑛4 = 1 𝑛5 = 1 27 𝑃27(𝑡)

𝑛1 = 0 𝑛2 = 2 𝑛3 = 1 𝑛4 = 0 𝑛5 = 0 28 𝑃28(𝑡)

𝑛1 = 0 𝑛2 = 2 𝑛3 = 1 𝑛4 = 0 𝑛5 = 1 29 𝑃29(𝑡)

𝑛1 = 0 𝑛2 = 2 𝑛3 = 1 𝑛4 = 1 𝑛5 = 0 30 𝑃30(𝑡)

𝑛1 = 0 𝑛2 = 2 𝑛3 = 1 𝑛4 = 1 𝑛5 = 1 31 𝑃31(𝑡)

𝑛1 = 0 𝑛2 = 2 𝑛3 = 2 𝑛4 = 0 𝑛5 = 0 32 𝑃32(𝑡)

𝑛1 = 0 𝑛2 = 2 𝑛3 = 2 𝑛4 = 0 𝑛5 = 1 33 𝑃33(𝑡)

𝑛1 = 0 𝑛2 = 2 𝑛3 = 2 𝑛4 = 1 𝑛5 = 0 34 𝑃34(𝑡)

𝑛1 = 0 𝑛2 = 2 𝑛3 = 2 𝑛4 = 1 𝑛5 = 1 35 𝑃35(𝑡)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 0 𝑛4 = 0 𝑛5 = 0 36 𝑃36(𝑡)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 0 𝑛4 = 0 𝑛5 = 1 37 𝑃37(𝑡)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 0 𝑛4 = 1 𝑛5 = 0 38 𝑃38(𝑡)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 0 𝑛4 = 1 𝑛5 = 1 39 𝑃39(𝑡)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 1 𝑛4 = 0 𝑛5 = 0 40 𝑃40(𝑡)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 1 𝑛4 = 0 𝑛5 = 1 41 𝑃41(𝑡)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 1 𝑛4 = 1 𝑛5 = 0 42 𝑃42(𝑡)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 1 𝑛4 = 1 𝑛5 = 1 43 𝑃43(𝑡)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 2 𝑛4 = 0 𝑛5 = 0 44 𝑃44(𝑡)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 2 𝑛4 = 0 𝑛5 = 1 45 𝑃45(𝑡)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 2 𝑛4 = 1 𝑛5 = 0 46 𝑃46(𝑡)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 2 𝑛4 = 1 𝑛5 = 1 47 𝑃47(𝑡)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 0 𝑛4 = 0 𝑛5 = 0 48 𝑃48(𝑡)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 0 𝑛4 = 0 𝑛5 = 1 49 𝑃49(𝑡)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 0 𝑛4 = 1 𝑛5 = 0 50 𝑃50(𝑡)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 0 𝑛4 = 1 𝑛5 = 1 51 𝑃51(𝑡)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 1 𝑛4 = 0 𝑛5 = 0 52 𝑃52(𝑡)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 1 𝑛4 = 0 𝑛5 = 1 53 𝑃53(𝑡)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 1 𝑛4 = 1 𝑛5 = 0 54 𝑃54(𝑡)

Page 173: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-157

𝑛1 = 1 𝑛2 = 1 𝑛3 = 1 𝑛4 = 1 𝑛5 = 1 55 𝑃55(𝑡)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 2 𝑛4 = 0 𝑛5 = 0 56 𝑃56(𝑡)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 2 𝑛4 = 0 𝑛5 = 1 57 𝑃57(𝑡)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 2 𝑛4 = 1 𝑛5 = 0 58 𝑃58(𝑡)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 2 𝑛4 = 1 𝑛5 = 1 59 𝑃59(𝑡)

𝑛1 = 1 𝑛2 = 2 𝑛3 = 0 𝑛4 = 0 𝑛5 = 0 60 𝑃60(𝑡)

𝑛1 = 1 𝑛2 = 2 𝑛3 = 0 𝑛4 = 0 𝑛5 = 1 61 𝑃61(𝑡)

𝑛1 = 1 𝑛2 = 2 𝑛3 = 0 𝑛4 = 1 𝑛5 = 0 62 𝑃62(𝑡)

𝑛1 = 1 𝑛2 = 2 𝑛3 = 0 𝑛4 = 1 𝑛5 = 1 63 𝑃63(𝑡)

𝑛1 = 1 𝑛2 = 2 𝑛3 = 1 𝑛4 = 0 𝑛5 = 0 64 𝑃64(𝑡)

𝑛1 = 1 𝑛2 = 2 𝑛3 = 1 𝑛4 = 0 𝑛5 = 1 65 𝑃65(𝑡)

𝑛1 = 1 𝑛2 = 2 𝑛3 = 1 𝑛4 = 1 𝑛5 = 0 66 𝑃66(𝑡)

𝑛1 = 1 𝑛2 = 2 𝑛3 = 1 𝑛4 = 1 𝑛5 = 1 67 𝑃67(𝑡)

𝑛1 = 1 𝑛2 = 2 𝑛3 = 2 𝑛4 = 0 𝑛5 = 0 68 𝑃68(𝑡)

𝑛1 = 1 𝑛2 = 2 𝑛3 = 2 𝑛4 = 0 𝑛5 = 1 69 𝑃69(𝑡)

𝑛1 = 1 𝑛2 = 2 𝑛3 = 2 𝑛4 = 1 𝑛5 = 0 70 𝑃70(𝑡)

𝑛1 = 1 𝑛2 = 2 𝑛3 = 2 𝑛4 = 1 𝑛5 = 1 71 𝑃71(𝑡)

Constructing a Markov state transition diagram become quite complex, challenging and

unmanageable due to the large number of system states. However, a systematic treatment of the

above enumerated system states can be performed using simple algorithm and implementing it in

tools such as MATLAB. The system states are organized systematically in layer fashion based on

the number of components failure (See Table 59 to Table 64). For instance, Layer 0 implies a

normal system states or zero (0) unit failure, Layer 1 implies a single (1) unit failure, Layer 2

implies a two (2) unit failure, and so on.

Page 174: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-158

Table 59: Systematic organization of system states in Layer fashion

Individual Unit State Combination System States

(𝑛) Drain unit

(𝑛1) Vent unit

(𝑛2) HX unit

(𝑛3) MK unit

(𝑛4) EF (𝑛5)

Layer 0 (Number of states with no failure= 1)

𝑛1 = 0 𝑛2 = 0 𝑛3 = 0 𝑛4 = 0 𝑛5 = 0 0 𝑃0(𝑡)

Layer 1 (Number of states with one= 7)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 0 𝑛4 = 0 𝑛5 = 0 36 𝑃36(𝑡)

𝑛1 = 0 𝑛2 = 1 𝑛3 = 0 𝑛4 = 0 𝑛5 = 0 12 𝑃12(𝑡)

⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮

Layer 2 (Number of states with two failures= 19)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 0 𝑛4 = 0 𝑛5 = 0 48 𝑃48(𝑡)

𝑛1 = 1 𝑛2 = 2 𝑛3 = 0 𝑛4 = 0 𝑛5 = 0 60 𝑃60(𝑡)

⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮

Layer 3 (Number of states with three failures= 25)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 1 𝑛4 = 0 𝑛5 = 0 52 𝑃52(𝑡)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 2 𝑛4 = 0 𝑛5 = 0 56 𝑃56(𝑡)

⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮

Layer 4 (Number of states with 4 failures= 16)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 1 𝑛4 = 1 𝑛5 = 0 54 𝑃54(𝑡)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 1 𝑛4 = 0 𝑛5 = 1 53 𝑃53(𝑡)

⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮

Layer 5 (Number of states with five failures= 4)

𝑛1 = 1 𝑛2 = 1 𝑛3 = 1 𝑛4 = 1 𝑛5 = 1 55 𝑃55(𝑡)

⋮ ⋮ ⋮ ⋮ ⋮ ⋮ ⋮

The possible number of transition from one layer to other can be determined systematically. Of

course, since no repair of the units are considered, the layer can only progress in an increasing

Page 175: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-159

manner. For instance, the system can make a transition from layer 2 to layer 3, but not vice versa.

The number of transition from each layer with their respective failure rates are given in Table 60.

Table 60: Transition from Layer 0 to Layer 1

From Transition rate To No.

𝑛 = 0:

{(𝑛1 = 0)(𝑛2 = 0)(𝑛3 = 0)

(𝑛4 = 0)(𝑛5 = 0)}

𝜆1𝐷𝑈

𝑛 = 36 1

𝜆1𝑉𝑈

𝑛 = 12 2

𝜆2𝑉𝑈

𝑛 = 24 3

𝜆1𝐻𝑋𝑈

𝑛 = 4 4

𝜆2𝐻𝑋𝑈

𝑛 = 8 5

𝜆1𝑀𝑊𝑈

𝑛 = 2 6

𝜆1𝐸𝐹

𝑛 = 1 7

Table 61: Transition from Layer 1 to Layer 2

From Transition rate To No.

𝑛 = 36:

{(𝑛1 = 1) (𝑛2 = 0) (𝑛3 =0) (𝑛4 = 0) (𝑛5 = 0)}

𝜆1𝑉𝑈

𝑛 = 48 1

𝜆2𝑉𝑈

𝑛 = 60 2

𝜆1𝐻𝑋𝑈

𝑛 = 40 3

𝜆2𝐻𝑋𝑈

𝑛 = 44 4

𝜆1𝑀𝑊𝑈

𝑛 = 38 5

𝜆1𝐸𝐹

𝑛 = 37 6

⋮ ⋮ ⋮ ⋮

Table 62: Transition from Layer 2 to Layer 3

From Transition rate To No.

𝑛 = 48:

{(𝑛1 = 1) (𝑛2 = 1) (𝑛3 = 0)

(𝑛4 = 0) (𝑛5 = 0)}

𝜆1𝐻𝑋𝑈

𝑛 = 52 1

𝜆2𝐻𝑋𝑈

𝑛 = 56 2

𝜆1𝑀𝑊𝑈

𝑛 = 50 3

𝜆1𝐸𝐹

𝑛 = 49 4

⋮ ⋮ ⋮ ⋮

Page 176: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-160

Table 63: Transition from Layer 3 to Layer 4

From Transition rate To No.

𝑛 = 52: {(𝑛1 = 1) (𝑛2 = 1) (𝑛3 = 1) (𝑛4 = 0) (𝑛5 = 0)}

𝜆1𝑀𝑊𝑈

𝑛 = 54 1

𝜆1𝐸𝐹

𝑛 = 53 2

⋮ ⋮ ⋮ ⋮

Table 64: Transition from Layer 4 to Layer 5

From Transition rate To No.

𝑛 = 54:

{(𝑛1 = 1). (𝑛2 = 1). (𝑛3 = 1) (𝑛4 = 1).(𝑛5 = 0)} 𝜆1

𝐸𝐹 𝑛 = 55 1

⋮ ⋮ ⋮ ⋮

The resulting ODEs from the state transition diagram is presented below in a layer fashion.

Layer 0:

𝑑𝑃0(𝑡)

𝑑𝑡= −(𝜆1

𝐷𝑈 + 𝜆1𝑉𝑈 + 𝜆2

𝑉𝑈 + 𝜆1𝐻𝑋𝑈 + 𝜆2

𝐻𝑋𝑈 + 𝜆1𝑀𝑊𝑈 + 𝜆1

𝐸𝐹)𝑃0(𝑡) (5-21)

Layer 1:

𝑑𝑃1(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃0(𝑡) − (𝜆1𝑀𝑊𝑈 + 𝜆1

𝐷𝑈 + 𝜆1𝑉𝑈 + 𝜆2

𝑉𝑈 + 𝜆1𝐻𝑋𝑈 + 𝜆2

𝐻𝑋𝑈)𝑃1(𝑡)

𝑑𝑃2(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃0(𝑡) − (𝜆1𝐸𝐹 + 𝜆1

𝐷𝑈 + 𝜆1𝑉𝑈 + 𝜆2

𝑉𝑈 + 𝜆1𝐻𝑋𝑈 + 𝜆2

𝐻𝑋𝑈)𝑃2(𝑡)

𝑑𝑃36(𝑡)

𝑑𝑡= 𝜆1

𝐷𝑈𝑃0(𝑡) − (𝜆1𝐸𝐹 + 𝜆1

𝑀𝑊𝑈 + 𝜆1𝑉𝑈 + 𝜆2

𝑉𝑈 + 𝜆1𝐻𝑋𝑈 + 𝜆2

𝐻𝑋𝑈)𝑃36(𝑡)

𝑑𝑃12(𝑡)

𝑑𝑡= 𝜆1

𝑉𝑈𝑃0(𝑡) − (𝜆1𝐸𝐹 + 𝜆1

𝑀𝑊𝑈 + 𝜆1𝐷𝑈 + 𝜆1

𝐻𝑋𝑈 + 𝜆2𝐻𝑋𝑈)𝑃12(𝑡)

𝑑𝑃24(𝑡)

𝑑𝑡= 𝜆2

𝑉𝑈𝑃0(𝑡) − (𝜆1𝐸𝐹 + 𝜆1

𝑀𝑊𝑈 + 𝜆1𝐷𝑈 + 𝜆1

𝐻𝑋𝑈 + 𝜆2𝐻𝑋𝑈)𝑃24(𝑡)

𝑑𝑃4(𝑡)

𝑑𝑡= 𝜆1

𝐻𝑋𝑈𝑃0(𝑡) − (𝜆1𝐸𝐹 + 𝜆1

𝑀𝑊𝑈 + 𝜆1𝐷𝑈 + 𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈)𝑃4(𝑡)

𝑑𝑃8(𝑡)

𝑑𝑡= 𝜆2

𝐻𝑋𝑈𝑃0(𝑡) − (𝜆1𝐸𝐹 + 𝜆1

𝑀𝑊𝑈 + 𝜆1𝐷𝑈 + 𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈 )𝑃8(𝑡)

(5-22)

Layer 2:

𝑑𝑃48(𝑡)

𝑑𝑡= 𝜆1

𝑉𝑈𝑃36(𝑡) + 𝜆1𝐷𝑈𝑃12(𝑡) − (𝜆1

𝐻𝑋𝑈 + 𝜆2𝐻𝑋𝑈 + 𝜆1

𝐸𝐹 + 𝜆1𝑀𝑊𝑈)𝑃48(𝑡)

Page 177: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-161

𝑑𝑃60(𝑡)

𝑑𝑡= 𝜆2

𝑉𝑈𝑃36(𝑡) + 𝜆1𝐷𝑈𝑃24(𝑡) − (𝜆1

𝐻𝑋𝑈 + 𝜆2𝐻𝑋𝑈 + 𝜆1

𝐸𝐹 + 𝜆1𝑀𝑊𝑈)𝑃60(𝑡)

𝑑𝑃40(𝑡)

𝑑𝑡= 𝜆1

𝐻𝑋𝑈𝑃36(𝑡) + 𝜆1𝐷𝑈𝑃4(𝑡) − (𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈 + 𝜆1

𝐸𝐹 + 𝜆1𝑀𝑊𝑈)𝑃40(𝑡)

𝑑𝑃44(𝑡)

𝑑𝑡= 𝜆2

𝐻𝑋𝑈𝑃36(𝑡) + 𝜆1𝐷𝑈𝑃8(𝑡) − (𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈 + 𝜆1

𝐸𝐹 + 𝜆1𝑀𝑊𝑈)𝑃44(𝑡)

𝑑𝑃38(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃36(𝑡) + 𝜆1𝐷𝑈𝑃2(𝑡) − (𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈 + 𝜆1

𝐸𝐹 + 𝜆1𝐻𝑋𝑈 + 𝜆2

𝐻𝑋𝑈)𝑃38(𝑡)

𝑑𝑃37(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃36(𝑡) + 𝜆1𝐷𝑈𝑃1(𝑡) − (𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈 + 𝜆1

𝑀𝑊𝑈 + 𝜆1𝐻𝑋𝑈 + 𝜆2

𝐻𝑋𝑈)𝑃37(𝑡)

𝑑𝑃16(𝑡)

𝑑𝑡= 𝜆1

𝐻𝑋𝑈𝑃12(𝑡) + 𝜆1𝑉𝑈𝑃4(𝑡) − (𝜆1

𝐷𝑈 + 𝜆1𝑀𝑊𝑈 + 𝜆1

𝐸𝐹)𝑃16(𝑡)

𝑑𝑃20(𝑡)

𝑑𝑡= 𝜆2

𝐻𝑋𝑈𝑃12(𝑡) + 𝜆1𝑉𝑈𝑃8(𝑡) − (𝜆1

𝐷𝑈 + 𝜆1𝑀𝑊𝑈 + 𝜆1

𝐸𝐹)𝑃20(𝑡)

𝑑𝑃14(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃12(𝑡) + 𝜆1𝑉𝑈𝑃2(𝑡) − (𝜆1

𝐷𝑈 + 𝜆1𝐸𝐹 + 𝜆1

𝐻𝑋𝑈 + 𝜆2𝐻𝑋𝑈)𝑃14(𝑡)

𝑑𝑃13(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃12(𝑡) + 𝜆1𝑉𝑈𝑃1(𝑡) − (𝜆1

𝐷𝑈 + 𝜆1𝑀𝑊𝑈 + 𝜆1

𝐻𝑋𝑈 + 𝜆2𝐻𝑋𝑈)𝑃13(𝑡)

𝑑𝑃28(𝑡)

𝑑𝑡= 𝜆1

𝐻𝑋𝑈𝑃24(𝑡) + 𝜆2𝑉𝑈𝑃4(𝑡) − (𝜆1

𝐷𝑈 + 𝜆1𝑀𝑊𝑈 + 𝜆1

𝐸𝐹)𝑃28(𝑡)

𝑑𝑃32(𝑡)

𝑑𝑡= 𝜆2

𝐻𝑋𝑈𝑃24(𝑡) + 𝜆2𝑉𝑈𝑃8(𝑡) − (𝜆1

𝐷𝑈 + 𝜆1𝑀𝑊𝑈 + 𝜆1

𝐸𝐹)𝑃32(𝑡)

𝑑𝑃26(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃24(𝑡) + 𝜆2𝑉𝑈𝑃2(𝑡) − (𝜆1

𝐷𝑈 + 𝜆1𝐸𝐹 + 𝜆1

𝐻𝑋𝑈 + 𝜆2𝐻𝑋𝑈)𝑃26(𝑡)

𝑑𝑃25(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃24(𝑡) + 𝜆2𝑉𝑈𝑃1(𝑡) − (𝜆1

𝐷𝑈 + 𝜆1𝑀𝑊𝑈 + 𝜆1

𝐻𝑋𝑈 + 𝜆2𝐻𝑋𝑈)𝑃25(𝑡)

𝑑𝑃6(𝑡)

𝑑𝑡= 𝜆1

𝐻𝑋𝑈𝑃2(𝑡) + 𝜆1𝑀𝑊𝑈𝑃4(𝑡) − (𝜆1

𝐷𝑈 + 𝜆1𝐸𝐹 + 𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈)𝑃6(𝑡)

𝑑𝑃5(𝑡)

𝑑𝑡= 𝜆1

𝐻𝑋𝑈𝑃1(𝑡) + 𝜆1𝐸𝐹𝑃4(𝑡) − (𝜆1

𝐷𝑈 + 𝜆1𝑀𝑊𝑈 + 𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈)𝑃5(𝑡)

𝑑𝑃10(𝑡)

𝑑𝑡= 𝜆2

𝐻𝑋𝑈𝑃2(𝑡) + 𝜆1𝑀𝑊𝑈𝑃8(𝑡) − (𝜆1

𝐷𝑈 + 𝜆1𝐸𝐹 + 𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈)𝑃10(𝑡)

𝑑𝑃9(𝑡)

𝑑𝑡= 𝜆2

𝐻𝑋𝑈𝑃1(𝑡) + 𝜆1𝐸𝐹𝑃8(𝑡) − (𝜆1

𝐷𝑈 + 𝜆1𝑀𝑊𝑈 + 𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈)𝑃9(𝑡)

𝑑𝑃3(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃1(𝑡) + 𝜆1𝐸𝐹𝑃2(𝑡) − (𝜆1

𝐷𝑈 + 𝜆1𝑉𝑈 + 𝜆2

𝑉𝑈 + 𝜆1𝐻𝑋𝑈 + 𝜆2

𝐻𝑋𝑈 )𝑃3(𝑡)

(5-23)

Page 178: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-162

Layer 3:

𝑑𝑃52(𝑡)

𝑑𝑡= 𝜆1

𝐻𝑋𝑈𝑃48(𝑡) + 𝜆1𝑉𝑈𝑃40(𝑡) + 𝜆1

𝐷𝑈𝑃16(𝑡) − (𝜆1𝐸𝐹 + 𝜆1

𝑀𝑊𝑈)𝑃52(𝑡)

𝑑𝑃56(𝑡)

𝑑𝑡= 𝜆2

𝐻𝑋𝑈𝑃48(𝑡) + 𝜆1𝑉𝑈𝑃44(𝑡) + 𝜆1

𝐷𝑈𝑃20(𝑡) − (𝜆1𝐸𝐹 + 𝜆1

𝑀𝑊𝑈)𝑃56(𝑡)

𝑑𝑃50(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃48(𝑡) + 𝜆1𝑉𝑈𝑃38(𝑡) + 𝜆1

𝐷𝑈𝑃14(𝑡) − (𝜆1𝐻𝑋𝑈 + 𝜆2

𝐻𝑋𝑈 + 𝜆1𝐸𝐹)𝑃50(𝑡)

𝑑𝑃49(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃48(𝑡) + 𝜆1𝑉𝑈𝑃37(𝑡) + 𝜆1

𝐷𝑈𝑃13(𝑡) − (𝜆1𝐻𝑋𝑈 + 𝜆2

𝐻𝑋𝑈 + 𝜆1𝑀𝑊𝑈)𝑃49(𝑡)

𝑑𝑃64(𝑡)

𝑑𝑡= 𝜆1

𝐻𝑋𝑈𝑃60(𝑡) + 𝜆2𝑉𝑈𝑃40(𝑡) + 𝜆1

𝐷𝑈𝑃28(𝑡) − (𝜆1𝑀𝑊𝑈 + 𝜆1

𝐸𝐹)𝑃64(𝑡)

𝑑𝑃68(𝑡)

𝑑𝑡= 𝜆2

𝐻𝑋𝑈𝑃60(𝑡) + 𝜆2𝑉𝑈𝑃44(𝑡) + 𝜆1

𝐷𝑈𝑃32(𝑡) − (𝜆1𝑀𝑊𝑈 + 𝜆1

𝐸𝐹)𝑃68(𝑡)

𝑑𝑃62(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃60(𝑡) + 𝜆2𝑉𝑈𝑃38(𝑡) + 𝜆1

𝐷𝑈𝑃26(𝑡) − (𝜆1𝐻𝑋𝑈 + 𝜆2

𝐻𝑋𝑈 + 𝜆1𝐸𝐹)𝑃62(𝑡)

𝑑𝑃61(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃60(𝑡) + 𝜆2𝑉𝑈𝑃37(𝑡) + 𝜆1

𝐷𝑈𝑃25(𝑡) − (𝜆1𝐻𝑋𝑈 + 𝜆2

𝐻𝑋𝑈 + 𝜆1𝑀𝑊𝑈)𝑃61(𝑡)

𝑑𝑃42(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃40(𝑡) + 𝜆1𝐻𝑋𝑈𝑃38(𝑡) + 𝜆1

𝐷𝑈𝑃6(𝑡) − (𝜆1𝐸𝐹 + 𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈)𝑃42(𝑡)

𝑑𝑃41(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃40(𝑡) + 𝜆1𝐻𝑋𝑈𝑃37(𝑡) + 𝜆1

𝐷𝑈𝑃5(𝑡) − (𝜆1𝑉𝑈 + 𝜆2

𝑉𝑈 + 𝜆1𝑀𝑊𝑈)𝑃41(𝑡)

𝑑𝑃46(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃44(𝑡) + 𝜆2𝐻𝑋𝑈𝑃38(𝑡) + 𝜆1

𝐷𝑈𝑃10(𝑡) − (𝜆1𝑉𝑈 + 𝜆2

𝑉𝑈 + 𝜆1𝐸𝐹)𝑃46(𝑡)

𝑑𝑃45(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃44(𝑡) + 𝜆2𝐻𝑋𝑈𝑃37(𝑡) + 𝜆1

𝐷𝑈𝑃9(𝑡) − (𝜆1𝑉𝑈 + 𝜆2

𝑉𝑈 + 𝜆1𝑀𝑊𝑈)𝑃45(𝑡)

𝑑𝑃39(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃38(𝑡) + 𝜆1𝑀𝑊𝑈𝑃37(𝑡) + 𝜆1

𝐷𝑈𝑃3(𝑡) − (𝜆1𝑉𝑈 + 𝜆2

𝑉𝑈 + 𝜆1𝐻𝑋𝑈 + 𝜆2

𝐻𝑋𝑈)𝑃39(𝑡)

𝑑𝑃18(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃16(𝑡) + 𝜆1𝐻𝑋𝑈𝑃14(𝑡) + 𝜆1

𝑉𝑈𝑃6(𝑡) − (𝜆1𝐷𝑈 + 𝜆1

𝐸𝐹)𝑃18(𝑡)

𝑑𝑃17(𝑡)

𝑑𝑡= 𝜆1

𝐻𝑋𝑈𝑃13(𝑡) + 𝜆1𝐸𝐹𝑃16(𝑡) + 𝜆1

𝑉𝑈𝑃5(𝑡) − (𝜆1𝐷𝑈 + 𝜆1

𝑀𝑊𝑈)𝑃17(𝑡)

𝑑𝑃22(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃20(𝑡) + 𝜆2𝐻𝑋𝑈𝑃14(𝑡) + 𝜆1

𝑉𝑈𝑃10(𝑡) − (𝜆1𝐷𝑈 + 𝜆1

𝐸𝐹)𝑃22(𝑡)

𝑑𝑃21(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃20(𝑡) + 𝜆2𝐻𝑋𝑈𝑃13(𝑡) + 𝜆1

𝑉𝑈𝑃9(𝑡) − (𝜆1𝐷𝑈 + 𝜆1

𝑀𝑊𝑈)𝑃21(𝑡)

𝑑𝑃15(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃14(𝑡) + 𝜆1𝑀𝑊𝑈𝑃13(𝑡) + 𝜆1

𝑉𝑈𝑃3(𝑡) − (𝜆1𝐷𝑈 + 𝜆1

𝐻𝑋𝑈 + 𝜆2𝐻𝑋𝑈)𝑃15(𝑡)

Page 179: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-163

𝑑𝑃30(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃28(𝑡) + 𝜆1𝐻𝑋𝑈𝑃26(𝑡) + 𝜆2

𝑉𝑈𝑃6(𝑡) − (𝜆1𝐷𝑈 + 𝜆1

𝐸𝐹)𝑃30(𝑡)

𝑑𝑃29(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃28(𝑡) + 𝜆1𝐻𝑋𝑈𝑃25(𝑡) + 𝜆2

𝑉𝑈𝑃5(𝑡) − (𝜆1𝐷𝑈 + 𝜆1

𝑀𝑊𝑈)𝑃29(𝑡)

𝑑𝑃34(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃32(𝑡) + 𝜆2𝐻𝑋𝑈𝑃26(𝑡) + 𝜆2

𝑉𝑈𝑃10(𝑡) − (𝜆1𝐷𝑈 + 𝜆1

𝐸𝐹)𝑃34(𝑡)

𝑑𝑃27(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃26(𝑡) + 𝜆1𝑀𝑊𝑈𝑃25(𝑡) + 𝜆2

𝑉𝑈𝑃3(𝑡) − (𝜆1𝐷𝑈 + 𝜆1

𝐻𝑋𝑈 + 𝜆2𝐻𝑋𝑈)𝑃27(𝑡)

𝑑𝑃33(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃32(𝑡) + 𝜆2𝐻𝑋𝑈𝑃25(𝑡) + 𝜆2

𝑉𝑈𝑃9(𝑡) − (𝜆1𝐷𝑈 + 𝜆1

𝑀𝑊𝑈)𝑃33(𝑡)

𝑑𝑃7(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃6(𝑡) + 𝜆1𝑀𝑊𝑈𝑃5(𝑡) + 𝜆1

𝐻𝑋𝑈𝑃3(𝑡) − (𝜆1𝐷𝑈 + 𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈)𝑃7(𝑡)

𝑑𝑃11(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃10(𝑡) + 𝜆1𝑀𝑊𝑈𝑃9(𝑡) + 𝜆2

𝐻𝑋𝑈𝑃3(𝑡) − (𝜆1𝐷𝑈 + 𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈)𝑃11(𝑡)

(5-24)

Layer 4:

𝑑𝑃54(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃52(𝑡) + 𝜆1𝐻𝑋𝑈𝑃50(𝑡) + 𝜆1

𝑉𝑈𝑃42(𝑡) + 𝜆1𝐷𝑈𝑃18(𝑡) − 𝜆1

𝐸𝐹𝑃54(𝑡)

𝑑𝑃53(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃52(𝑡) + 𝜆1𝐻𝑋𝑈𝑃49(𝑡) + 𝜆1

𝑉𝑈𝑃41(𝑡) + 𝜆1𝐷𝑈𝑃17(𝑡) − 𝜆1

𝑀𝑊𝑈𝑃53(𝑡)

𝑑𝑃58(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃56(𝑡) + 𝜆2𝐻𝑋𝑈𝑃50(𝑡) + 𝜆1

𝑉𝑈𝑃46(𝑡) + 𝜆1𝐷𝑈𝑃22(𝑡) − 𝜆1

𝐸𝐹𝑃58(𝑡)

𝑑𝑃57(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃56(𝑡) + 𝜆2𝐻𝑋𝑈𝑃49(𝑡) + 𝜆1

𝑉𝑈𝑃45(𝑡) + 𝜆1𝐷𝑈𝑃21(𝑡) − 𝜆1

𝑀𝑊𝑈𝑃57(𝑡)

𝑑𝑃51(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃50(𝑡) + 𝜆1𝑀𝑊𝑈𝑃49(𝑡) + 𝜆1

𝑉𝑈𝑃39(𝑡) + 𝜆1𝐷𝑈𝑃15(𝑡) − (𝜆1

𝐻𝑋𝑈 + 𝜆2𝐻𝑋𝑈)𝑃51(𝑡)

𝑑𝑃66(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃64(𝑡) + 𝜆1𝐻𝑋𝑈𝑃62(𝑡) + +𝜆2

𝑉𝑈𝑃42(𝑡) + 𝜆1𝐷𝑈𝑃30(𝑡) − 𝜆1

𝐸𝐹𝑃66(𝑡)

𝑑𝑃65(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃64(𝑡) + 𝜆1𝐻𝑋𝑈𝑃61(𝑡) + +𝜆2

𝑉𝑈𝑃41(𝑡) + 𝜆1𝐷𝑈𝑃29(𝑡) − 𝜆1

𝑀𝑊𝑈𝑃65(𝑡)

𝑑𝑃70(𝑡)

𝑑𝑡= 𝜆1

𝑀𝑊𝑈𝑃68(𝑡) + 𝜆2𝐻𝑋𝑈𝑃62(𝑡) + +𝜆2

𝑉𝑈𝑃46(𝑡) + 𝜆1𝐷𝑈𝑃34(𝑡) − 𝜆1

𝐸𝐹𝑃70(𝑡)

𝑑𝑃69(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃68(𝑡) + 𝜆2𝐻𝑋𝑈𝑃61(𝑡) + +𝜆2

𝑉𝑈𝑃45(𝑡) + 𝜆1𝐷𝑈𝑃33(𝑡) − 𝜆1

𝑀𝑊𝑈𝑃69(𝑡)

𝑑𝑃63(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃62(𝑡) + 𝜆1𝑀𝑊𝑈𝑃61(𝑡) + 𝜆2

𝑉𝑈𝑃39(𝑡) + 𝜆1𝐷𝑈𝑃27(𝑡) − (𝜆1

𝐻𝑋𝑈 + 𝜆2𝐻𝑋𝑈)𝑃63(𝑡)

Page 180: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-164

𝑑𝑃43(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃42(𝑡) + 𝜆1𝑀𝑊𝑈𝑃41(𝑡) + 𝜆1

𝐻𝑋𝑈𝑃39(𝑡) + 𝜆1𝐷𝑈𝑃7(𝑡) − (𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈)𝑃43(𝑡)

𝑑𝑃47(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃46(𝑡) + 𝜆1𝑀𝑊𝑈𝑃45(𝑡) + 𝜆2

𝐻𝑋𝑈𝑃39(𝑡) + 𝜆1𝐷𝑈𝑃11(𝑡) − (𝜆1

𝑉𝑈 + 𝜆2𝑉𝑈)𝑃47(𝑡)

𝑑𝑃19(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃18(𝑡) + 𝜆1𝑀𝑊𝑈𝑃17(𝑡) + 𝜆1

𝐻𝑋𝑈𝑃15(𝑡) + 𝜆1𝑉𝑈𝑃7(𝑡) − 𝜆1

𝐷𝑈𝑃19(𝑡)

𝑑𝑃23(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃22(𝑡) + 𝜆1𝑀𝑊𝑈𝑃21(𝑡) + 𝜆2

𝐻𝑋𝑈𝑃15(𝑡) + 𝜆1𝑉𝑈𝑃11(𝑡) − 𝜆1

𝐷𝑈𝑃23(𝑡)

𝑑𝑃31(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃30(𝑡) + 𝜆1𝑀𝑊𝑈𝑃29(𝑡) + 𝜆1

𝐻𝑋𝑈𝑃27(𝑡) + 𝜆2𝑉𝑈𝑃7(𝑡) − 𝜆1

𝐷𝑈𝑃31(𝑡)

𝑑𝑃35(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃34(𝑡) + 𝜆1𝑀𝑊𝑈𝑃33(𝑡) + 𝜆2

𝐻𝑋𝑈𝑃27(𝑡) + 𝜆2𝑉𝑈𝑃11(𝑡) − 𝜆1

𝐷𝑈𝑃35(𝑡)

(5-25)

Layer 5:

𝑑𝑃55(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃54(𝑡) + 𝜆1𝑀𝑊𝑈𝑃53(𝑡) + 𝜆1

𝐻𝑋𝑈𝑃51(𝑡) + 𝜆1𝑉𝑈𝑃43(𝑡) + 𝜆1

𝐷𝑈𝑃19(𝑡) − 𝜆1𝐻𝑋𝑈𝑃55(𝑡)

𝑑𝑃59(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃58(𝑡) + 𝜆1𝑀𝑊𝑈𝑃57(𝑡) + 𝜆2

𝐻𝑋𝑈𝑃51(𝑡) + 𝜆1𝑉𝑈𝑃47(𝑡) + 𝜆1

𝐷𝑈𝑃23(𝑡) + 𝜆1𝐻𝑋𝑈𝑃55(𝑡)

𝑑𝑃67(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃66(𝑡) + 𝜆1𝑀𝑊𝑈𝑃65(𝑡) + 𝜆1

𝐻𝑋𝑈𝑃63(𝑡) + 𝜆2𝑉𝑈𝑃43(𝑡) + 𝜆1

𝐷𝑈𝑃31(𝑡) − 𝜆1𝐻𝑋𝑈𝑃67(𝑡)

𝑑𝑃71(𝑡)

𝑑𝑡= 𝜆1

𝐸𝐹𝑃70(𝑡) + 𝜆1𝑀𝑊𝑈𝑃69(𝑡) + 𝜆2

𝐻𝑋𝑈𝑃63(𝑡) + 𝜆2𝑉𝑈𝑃47(𝑡) + 𝜆1

𝐷𝑈𝑃35(𝑡) + 𝜆1𝐻𝑋𝑈𝑃67(𝑡)

(5-26)

The above coupled ODEs are solved using finite element method in the computer code Fortran95.

The following initial conditions at 𝑡 = 0 is assumed for system analysis:

𝑃𝑛(𝑡 = 0) = 1; 𝑓𝑜𝑟 𝑛 = 00; 𝑓𝑜𝑟 𝑛 ≠ 0

(5-27)

Sample results obtained from the code is presented in Table 65. Only three (3) system state

transition probabilities with 10 time steps are presented for comprehension. Moreover, due to the

volume of information, the results depiction significantly requires a large number of pages

(approximately 30 pages). Sample transition state probabilities is provided in Appendix V. It is

obvious from the result that the probability of the system state 𝑛 = 0 (all units in normal state)

decreases as time progresses, i.e., the failure probability of the unit increases with time, which is

intuitive. From a total of 72 systems states, it was found that the states 𝑃1(𝑡), 𝑃24(𝑡), 𝑃36(𝑡), 𝑃4(𝑡),

Page 181: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-165

𝑃8(𝑡), 𝑃37(𝑡), 𝑃25(𝑡) and 𝑃60(𝑡) are most likely to occur with the state probabilities within the

range 10−03 and 10−06 for 𝑘 = 9. Of course, the simulation time steps could be increased,

however it requires significant amount of time to obtain the solution from the computer code.

Among the above-mentioned states, 𝑃1(𝑡) have the highest probability of failure. This implies that

the ICS system is likely to fail due to an envelope failure or the pressure boundary being

compromised, i.e., a pipe break or rupture as compared to the other system states. This scenario

directly challenges the IC loop, and a failure of which could rapidly result to a total system failure.

A failure of the ICS to maintain its loop pressure will directly affect the reactor coolant system

pressure, which can result in a operation of the safety relief valves. System state 𝑃1(𝑡) is followed

by the states 𝑃24(𝑡) and 𝑃36(𝑡) which are the vent unit and drain unit failure respectively. An

occurrence of any of these two states will result to a system failure. For instance, the failure of

vent unit to vent non-condensable gases from the IC loop will significantly degrade the heat

transfer rate and hence the ICS performance. It is evident that a failure of drain unit will directly

result to a system failure since the IC loop will be degraded and the condensate coolant flow path

will be affected. The reason for 𝑃1(𝑡) being higher than 𝑃24(𝑡) and 𝑃36(𝑡) is due to the redundancy

provided for the vent and condensate return unit. Whereas, 𝑃36(𝑡) being lower than 𝑃24(𝑡) is due

to the diversity provided in the condensate return valves as compared to the identical vent valves.

Regardless, any of the above scenarios will have a direct influence on the IC loop pressure and

coolant flow, which in turn will affect the reactor coolant system pressure. The predicted time-

dependent stochastic ICS system behavior can be utilized in the next step of the project to couple

the model with deterministic system evolution, thus enabling an analyst to predict an integrated

determinist and probabilistic system evolution.

Page 182: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-166

Table 65: Sample ICS state transition probabilities with 10 time steps

1. For system state 𝑛 = 0:

𝑘 0 1 2 3 4 5 6 7 8 9

𝑃0(𝑡) 1 0.9995 0.998 0.998 0.997 0.997 0.996 0.996 0.995 0.995

2. For system state 𝑛 = 1:

𝑘 0 1 2 3 4 5

𝑃1(𝑡) 0 2.74E-04 5.48E-04 8.21E-04 1.09E-03 1.37E-03

6 7 8 9

1.64E-03 1.91E-03 2.19E-03 2.46E-03

3. For system state 𝑛 = 2:

𝑘 0 1 2 3 4 5

𝑃2(𝑡) 0 3.21E-08 6.41E-08 9.61E-08 1.28E-07 1.591E-07

6 7 8 9

1.92E-07 2.24E-07 2.55E-07 2.87E-07

Note: State transition probabilities are systematically organized by system configuration. A sample total state transition probabilities for

5 time steps is given in Appendix V.

Page 183: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-167

CHAPTER 6: SUMMARY AND CONCLUSIONS

This chapter presents the conclusion of the research work and some recommendations for future

research. The conclusion is derived from all the chapters of the thesis. Recommendations for

potential future research activities as an extension of the present work as well as ongoing and near-

term activities to be performed for the CRP are also noted in this chapter.

Conclusions

This thesis investigates the application and comparison of classical and dynamic PSA techniques

that are utilized for safety analysis of nuclear power plants. The classical techniques include the

fault tree, event tree and classical Markov model, whereas the dynamic methodologies include the

Markov-CCMT model and Dynamic Flowgraph Method. The capabilities and limitations of the

techniques are demonstrated by applying it to a benchmark liquid level control system exhibiting

dynamic characteristics. The failure probability of the top events (drained and overflow) are

evaluated using each method, and the predicted results are compared. The system reliability

analysis using FT/ET are performed using the CAFTA code, whereas states transition probabilities

in classical Markov model are executed in Fortran95 code. The DFM model of the benchmark

system is developed using the code DYMONDA, and the coupling of Markov-CCMT model is

performed using Fortran95 and MATLAB tools. The contribution of this research includes;

• Demonstration of the advantages of dynamic techniques over classical methods for

reliability modelling and analysis of system exhibiting dynamic characteristics;

• Formulated the time-dependent reliability analysis of system with multiple failure modes

and multi-state components;

• Generation of timed fault trees from DFM model through the backtracking process that

can be integrated into an existing conventional PSA model;

• Performed a detailed comparison of the two dynamic PSA methods, i.e., DFM and

Markov-CCMT models with regard to time-dependent modelling capability, dynamic

accident sequence generation, modelling complexity and computational time required;

• Developed a roadmap for the Coordinated Research Project and a novel approach to

model and compute the functional reliability of passive safety systems.

Page 184: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-168

The classical techniques are static in nature except the time-dependent Markov model, and

consider an instant system failure for a given minimal cut-set. This leads one to conclude that given

the set of possible system failures, the classical approach is a systematic approach linking known

anticipated events. However, it has been shown in this research that a failure event does not

necessarily result in an instant system failure, rather there is a time delay which is dependent on

the evolution of the state variables. The knowledge of time delay is important since it may be the

time available for the operators to respond to an initiating event, or the time available for the next

frontline system to activate on demand. It is self-evident that classical approach limits itself to the

state of the hardware system, whereas the dynamic approach simultaneously considers both the

hardware states as well as the magnitude of the state variables. The top events in classical

techniques are defined only in terms of failure probability of the hardware state, whereas dynamic

methods define the top events in terms of magnitude of state variables. This research thus

demonstrates that the dynamic methods can capture the time element and the time-dependent

probabilistic accident sequence evolution, and hence provides a more detailed approach to

evaluating complex dynamic system behaviour.

However, this comes at the expense of model complexity, and computationally, the time required

to process large transition matrix. Additional factors include, the large number of simulation runs

and complicated coupling process/algorithm. For instance, consider the Markov-CCMT model.

This physical model of the liquid level control system was created using MATLAB and verified

via analytical method. The stochastic model was developed systematically via Markov transition

diagram, and the generated coupled ODEs was solved in Fortran95 code. State variable

discretization, i.e., five (5) computational cells was performed in the MATLAB environment.

Based on the number of computational cells and possible systems states from the stochastic model,

the number of runs to be performed was determined. The evolution of state variable in the state-

space for each run is mapped, and its conditional probabilities computed using the CCMT model.

The coupling of the two models i.e., Markov and CCMT, and evaluation of the transition

probability matrix was performed in MATLAB. The size of the transition matrix increases with an

increase in number of system states and computational cells, and hence evaluation of the matrix

become considerably complicated requiring significant computational time.

The impact of state ordering of the accident sequences is also demonstrated such that the dynamic

techniques model ordering of failures and thus provides more information of the failure sequence.

Page 185: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-169

DFM and Markov-CCMT model strictly account for failure ordering which is probabilistic in

nature and enable probabilistic dynamic system evolution, i.e., simultaneously accounting for

failure ordering as well as system dynamics. This is due to the fact that both system state variables

and hardware states are accounted for simultaneously. In contrast, FT technique represents a

system failure without any particular ordering index; that is, the sequence is fixed apriori by experts

undertaking ET analysis. Here, analysis has shown that DFM present advantages in this regard.

Furthermore, DFM can account for stochastic accident sequence evolution, in contrast to the

classical techniques and Markov-CCMT model where the possible system states are set a priori by

the analyst. For instance, with an increase in backtracking steps, basic events in the prime

implicants (analogous to minimal cut-sets in fault tree) increases, whereas the possible system state

in Markov-CCMT model remains the same except with a deviation in the failure probability. This

is of vital importance as the evolution of accident sequence can be observed from the individual

timed-prime implicants (time stamped). DFM however lacks the capability to model deviation in

component/system failure rate and demand frequency influenced by a prior failure event. The

predicted failure probability of the benchmark system for both overflow and drained scenarios

show significant deviation from classical and dynamic techniques. The difference in the predicted

failure probabilities is attributed to lack of treatment of dynamic interactions among the units

through the state variable, and consideration of independence among unit failure modes in the

classical approach. In contrast, dynamic methods account for competition within the failure modes

and treat all unit modes in an integrated fashion. The classical approach yields conservative results

and thus serves to identify logical correctness and establishes conservative relationships between

top events and system end states. The viability of integrating the results obtained from a DFM

model to an existing static PSA model is depicted by generating timed-fault trees through the

backtracking algorithm. we note that even though it is impractical to model the entire plant using

DFM, specific system important to safety can be modelled with DFM, and its results can be

integrated after post-processing of timed-prime implicants to construct timed-fault trees.

The current research also achieves its objective in developing a preliminary roadmap and selection

of a benchmark passive safety system for the coordinated research project. The passive ICS was

selected on the basis that: many current generation SMRs concepts implements ICS, availability

of research materials and the objective within the international community to determine the

functional reliability of system based on natural circulation. A dynamic and integrated approach

Page 186: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-170

for safety assessment of passive systems is developed and presented in Chapter 5. A Markov-

CCMT model was selected as the most viable candidate for ICS system modelling due to its nature

of tight coupling, and the capability to capture dynamic interactions between systems that results

to a deviation in demand frequency of the interacting systems. As a first step towards integrated

safety assessment, two classical approaches to reliability modelling of the passive ICS were

proposed: (a) small Markov model and large fault tree; and (2) small fault tree and large Markov

model. The first approach is largely classical (static in nature) and provide results in terms of cut-

sets and is unable to incorporate dynamic characteristics such as state variable evolution and state

ordering. The latter approach enables one to model system level time-dependent characteristics as

well as merge unit/component states using FT failure rate evaluation method. This results in a

reduced system state. This indicates that the latter approach for stochastic system modelling is

practical, and subsequently eventual coupling with ICS deterministic model. Reliability of the ICS

and hence the system state transition probabilities are predicted using the latter approach. The

analysis confirms a high system reliability for the required mission time.

Recommendations for Future Work

Based on the thesis research activities performed, recommendations and anticipated directions are

presented in this sub-section. The recommendations also include the activities to be performed for

the research project, and is aligned with the roadmap of the project outlined in Chapter 5. The

potential subject of research includes:

1. Development of a mechanized coupling process of stochastic and deterministic system

model (best estimate code) through a user-friendly computer code. For instance, the

coupling of stochastic model in Fortran95 and physical model in RELAP5 can be realized

using Python as the coupling tool.

2. Most, if not all, of the existing NPPs are based on static classical PSA approach. This

requires post-processing of results obtained from dynamic techniques in order to integrate

into an existing PSA model. Due to the significant volume of information obtained from

dynamic methods, post-processing is itself a challenge. This will require time and

investigation. Furthermore, there exist no standard technique for post-processing and is

thus, analyst dependent. Additional work should be performed to develop a systematic

approach for post-processing of results that can be integrated with an existing PSA model.

Page 187: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-171

3. The ICS is a passive reactor pressure depressurization system. However, in almost all

current generation reactors, an active system is also dedicated for the same purpose, such

as safety relief valves (SRVs) or automatic depressurization systems (ADS). The modelling

of complex dynamic interactions between the two-depressurization system through the

RCS pressure evolution needs further research. This will enable one to comprehend how

dynamic interactions can influence the failure rate, demand frequency and hence the

reliability of interacting systems. This may be particularly important for PSSs that function

on a small driving force which is highly sensitive to system parameters such as heat loss,

presence of non-condensable gases, pipe fouling and oxidation. Unlike the active systems

where these parameters are negligible due to the high driving force. These phenomenon in

turn influence the state variable (i.e., RCS pressure) evolution, and thus the demand

frequency and reliability of interacting systems.

In conclusion, risk assessment of NPPs requires accurate estimates of the frequency of accident

scenarios and adequate treatment of dependencies among multiple failure events. The ET/FT

methodology currently used is fundamentally well-suited to conservatively treat dependencies that

can be expressed in static and logical terms. However, it is not as well-suited to treat dependencies

associated with time-dependent processes or continuously varying variables. The methodology

does not explicitly carry information concerning the evolution of state variables and operator state

that may couple a number of failure events together. the three major severe accidents (TMI,

Chernobyl and Fukushima Daiichi) substantiate this observation. The study shows that the two

dynamic methods provide a promising pathway to bridging the gap and overcoming the limitations

encountered in classical PSA; hence providing risk-significant information and insights into the

probabilistic system dynamic evolution. Lastly, dynamic PSA techniques can be used as a

complement to the classical approach in that, it provides information on the system dynamic

evolution besides providing confirmatory information as that available via classical techniques.

Page 188: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-172

REFERENCES

1. Acosta, C., and N. Siu, “Dynamic event tree analysis method (DETAM) for accident

sequence analysis”, Cambridge, MA: Dept. of Nuclear Engineering, Massachusetts Institute

of Technology (1991).

2. Acosta, C., and N. Siu, "Dynamic event trees in accident sequence analysis: application to

steam generator tube rupture", Reliability Engineering & System Safety 41.2 (1993): 135-

154.

3. Aldemir, T., "Computer-Assisted Markov Failure Modeling of Process Control System," IEEE

Transactions on Reliability, 36 (1987):1, 133-144.

4. Aldemir, T., "A survey of dynamic methodologies for probabilistic safety assessment of

nuclear power plants", Annals of Nuclear Energy 52 (2013): 113-124.

5. Aldemir, T., D. Miller, M. Stovsky, J. Kirschenbaum, P. Bucci, A. Fentiman, L. Mangan, and

S. Arndt, "NUREG/CR 6901: Current state of reliability modeling methodologies for digital

systems and their acceptance criteria for nuclear power plant assessments", Office of Nuclear

Regulatory Research, US Nuclear Regulatory Commission, Washington DC (2006).

6. Aldemir, T., et al., "A benchmark implementation of two dynamic methodologies for the

reliability modeling of digital instrumentation and control systems", NUREG/CR-6985, US

Nuclear Regulatory Commission, Washington DC (2009).

7. Aldemir, T., et al., "Dynamic reliability modeling of digital instrumentation and control

systems for nuclear reactor probabilistic risk assessments", NUREG0CR-6942, US Nuclear

Regulatory Commission Washington DC (2007).

8. Amoia, V., G. D. Micheli, and M. Santomauro, "Computer-oriented formulation of transition-

rate matrices via Kronecker algebra", IEEE Transactions on Reliability 30.2 (1981): 123-132.

9. ASCA Inc., “Dymonda 7.0 Software Guide,” ASCA Inc., California, USA, 2013.

10. Basharin, G. P., Langville, A. N., and Naumov, V. A., "The life and work of AA Markov",

Linear algebra and its applications 386, 3-26, (2004).

11. Beeson, S. C., “Non-coherent Fault Tree Analysis”, PhD Diss., Loughborough University

(2002).

12. Belhadj, M., and T. Aldemir. "The Cell to Cell Mapping technique and Chapman-Kolmogorov

representation of system dynamics", Journal of Sound and Vibration 181.4 (1995): 687-707.

Page 189: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-173

13. Bucci, P., et al., "Construction of event-tree/fault-tree models from a Markov approach to

dynamic system reliability", Reliability Engineering & System Safety 93.11 (2008): 1616-

1627.

14. Burgazzi, L., "Open issues associated with passive safety systems reliability assessment",

International Conference on Opportunities and Challenges for Water Cooled Reactors in the

21st Century, Vienna (2009).

15. C. J. Garrett, S. B. Guarro and G. E. Apostolakis, “The Dynamic Flowgraph Methodology for

Assessing the Dependability of Embedded Software Systems,” IEEE Transactions on Systems,

Man, and Cybernetics, vol. 25, pp. 824-840 (1995).

16. Cafaro, G., F. Corsi, and F. Vacca, "Multistate Markov models and structural properties of the

transition-rate matrix", IEEE Transactions on Reliability 35.2 (1986): 192-200.

17. Canadian Nuclear Safety Commission, "REGDOC-2.4. 2, Safety Analysis: Probabilistic

Safety Assessment (PSA) for Nuclear Power Plants", Ottawa (2014).

18. Canadian Nuclear Safety Commission, "REGDOC-2.5. 2, Design of Reactor Facilities:

Nuclear Power Plants", Regulatory Document, Ottawa (2014).

19. Chen, Z., “Global Analysis, Control, and Bayesian Estimation of Nonlinear Dynamic Systems

Using Cell-to-cell Mapping”, PhD Dissertation, Cleveland State University (2004).

20. Devooght, J., and C. Smidts, "Probabilistic dynamics as a tool for dynamic PSA", Reliability

Engineering & System Safety 52.3 (1996): 185-196.

21. Devooght, J., and C. Smidts, "Probabilistic dynamics as a tool for dynamic PSA", Reliability

Engineering & System Safety 52.3 (1996): 185-196.

22. Devooght, J., and C. Smidts, "Probabilistic reactor dynamics—I: the theory of continuous

event trees", Nuclear Science and Engineering 111.3 (1992): 229-240.

23. Devooght, J., and C. Smidts. "Probabilistic reactor dynamics—I: the theory of continuous

event trees", Nuclear Science and Engineering 111.3 (1992a): 229-240.

24. Devooght, J., and C. Smidts. "Probabilistic reactor dynamics—III. A framework for time-

dependent interaction between operator and reactor during a transient involving human

error", Nuclear Science and Engineering 112.2 (1992b): 101-113.

25. Dinca, L. G., “A probabilistic approach to parameter estimation towards fault diagnosis in

nonlinear dynamic systems”, PhD Dissertation, The Ohio State University (1997).

Page 190: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-174

26. Electric Power Research Institute, “CAFTA Fault Tree Analysis System Version 6.0”, EPRI

Inc., Charlotte, NC, USA (2013).

27. Fussell, J. B., "A formal methodology for fault tree construction", Nuclear Science and

Engineering 52.4 (1973): 421-432.

28. Garrett, C. J., and G. E. Apostolakis, "Automated hazard analysis of digital control systems",

Reliability Engineering & System Safety 77.1 (2002): 1-17.

29. Garribba, S., E. Guagnini, and P. Mussio, "Multiple-valued logic trees: meaning and prime

implicants", IEEE Transactions on Reliability 34.5 (1985): 463-472.

30. GE Nuclear Energy, “SBWR Standard Safety Analysis”, MFN 04-138 (1992).

31. Gomes, I. B., P. L. C. Saldanha, and P. F. F. F. Melo, "A cell-to-cell Markovian model for the

reliability of a digital control system of a steam generator", Proceedings of the 2013

International Nuclear Atlantic Conference-INAC (2013).

32. Hsu, C. S., "A generalized theory of cell-to-cell mapping for nonlinear dynamical systems",

Journal of Applied Mechanics 48.3 (1981): 634-642.

33. Hsu, C. S., "A theory of cell-to-cell mapping dynamical systems", Journal of Applied

Mechanics 47.4 (1980a): 931-939.

34. Hsu, C. S., and R. S. Guttalu. "An unravelling algorithm for global analysis of dynamical

systems: An application of cell-to-cell mappings", Journal of Applied Mechanics 47.4 (1980b):

940-948.

35. International Atomic Energy Agency, “Probabilistic Safety Assessment”, INSAG-6, a report

by the International Nuclear Safety Advisory Group, IAEA, Vienna (1992).

36. Jafari, J., et al. "Reliability evaluation of a natural circulation system." Nuclear Engineering

and Design 224.1 (2003): 79-104.

37. Khan, H. J., and U. S. Rohatgi, “Performance characterization of isolation condenser of

SBWR”, BNL-NUREG-47960; CONF-921102-25, Brookhaven National Lab., NY (1992).

38. Kirschenbaum, J., et al. "A benchmark system for comparing reliability modeling approaches

for digital instrumentation and control systems", Nuclear Technology 165.1 (2009): 53-95.

39. Kirschenbaum, J., et al., "A benchmark system for comparing reliability modeling

approaches for digital instrumentation and control systems", Nuclear Technology 165.1

(2009): 53-95.

Page 191: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-175

40. Kumamoto, H. and Ernest J. H., “Probabilistic risk assessment and management for engineers

and scientists”, Wiley-IEEE (2000).

41. Kumamoto, H., “Satisfying safety goals by probabilistic risk assessment”, Springer Science

& Business Media (2007).

42. Labeau, P., C. Smidts, and S. Swaminathan, "Dynamic reliability: towards an integrated

platform for probabilistic risk assessment", Reliability Engineering & System Safety 68.3

(2000): 219-254.

43. Lesanovsky, A., "Multistate Markov models for systems with dependent units", IEEE

Transactions on Reliability 37.5 (1988): 505-511.

44. Lorenzo, G., et al., "Assessment of an isolation condenser of an integral reactor in view of

uncertainties in engineering parameters." Science and technology of Nuclear Installations

(2011).

45. M. Belhadj, M. Hassan and T. Aldemir, "On the need for dynamic methodologies in risk and

reliability studies", Reliability Engineering and System Safety, vol. 38, pp. 219-236, 1992.

46. M. Hassan and T. Aldemir, "A Data Base Oriented Dynamic Methodology for the Failure

Analysis of Closed Loop Control Systems in Process Plants", Reliability Engineering and

System Safety, vol. 27, pp. 275-322, 1990.

47. Mandelli, D., “Reliability Modeling of Digital Control Systems Using the Markov/cell-to-cell

Mapping Technique”, MS Dissertation, The Ohio State University (2008).

48. Marques, M., et al. "Methodology for the reliability evaluation of a passive system and its

integration into a probabilistic safety assessment", Nuclear Engineering and Design 235.24

(2005): 2612-2631.

49. Marques, M., et al. "Methodology for the reliability evaluation of a passive system and its

integration into a probabilistic safety assessment." Nuclear Engineering and Design 235.24

(2005): 2612-2631.

50. Marseguerra, M., and E. Zio, "Monte Carlo approach to PSA for dynamic process systems",

Reliability Engineering & System Safety 52.3 (1996): 227-241.

51. Marseguerra, M., et al., "A concept paper on dynamic reliability via Monte Carlo

simulation", Mathematics and Computers in simulation 47.2-5 (1998): 371-382.

Page 192: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-176

52. McNelles, P., “Dynamic safety assessment of FPGA-based safety critical systems with

applications in nuclear power generation”, PhD Dissertation, University of Ontario Institute of

Technology (2016).

53. Mezio, F., et al. "Integration of the functional reliability of two passive safety systems to

mitigate a SBLOCA+ BO in a CAREM-like reactor PSA." Nuclear Engineering and Design

270 (2014): 109-118.

54. Nayak, A. K., A. Chandrakar, and G. Vinod, "A review: passive system reliability analysis–

accomplishments and unresolved issues", Frontiers in Energy Research 2 (2014): 40.

55. Nayak, A. K., and R. K. Sinha. "Role of passive systems in advanced reactors." Progress in

Nuclear Energy 49.6 (2007): 486-498.

56. Nayak, A. K., et al. "Passive system reliability analysis using the APSRA methodology."

Nuclear Engineering and Design 238.6 (2008(a)): 1430-1440.

57. Nayak, A. K., et al. "Reliability assessment of passive containment isolation system using

APSRA methodology." Annals of Nuclear Energy 35.12 (2008(b)): 2270-2279.

58. Nieuwhof, G. W. E., "An introduction to fault tree analysis with emphasis on failure rate

evaluation", Microelectronics Reliability 14.2 (1975): 105-119.

59. Ogunbiyi, E. I., "Application of Decision Tables to Risk Analysis Studies", PhD Dissertation,

University of Houston, Texas (1981b).

60. Ogunbiyi, E. I., and E. J. Henley, "Irredundant forms and prime implicants of a function with

multistate variables", IEEE Transactions on Reliability 30.1 (1981a): 39-42.

61. Papazoglou, I. A., and E. P. Gyftopoulos, "Markov processes for reliability analyses of large

systems", IEEE Transactions on Reliability 26.3 (1977): 232-237.

62. Platz, O., "A Markov model for common-cause failures", Reliability Engineering 9.1 (1984):

25-31.

63. Privault, N., “Understanding Markov chains: examples and applications”, Springer Science &

Business Media (2013).

64. Pukite, P., and Pukite, J., “Modeling for Reliability Analysis: Markov Modeling for Reliability,

Maintainability, Safety, and Supportability Analyses of Complex Computer Systems”, Wiley-

IEEE Press (1998).

65. Ruijters, E., and M. Stoelinga. "Fault tree analysis: a survey of the state-of-the-art in

modeling, analysis and tools", Computer Science Review 15 (2015): 29-62.

Page 193: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-177

66. Guarro, S., M. Yau and M. Motamed, “Development of Tools for Safety Analysis of Control

Software in Advanced Reactors”, NUREG/CR-6465, U.S. Nuclear Regulatory Commission,

Washington DC (1996).

67. Salem, S. L., G. E. Apostolakis, and D. Okrent, "A new methodology for the computer-aided

construction of fault trees", Annals of Nuclear Energy 4.9-10 (1977): 417-433.

68. Salem, S. L., G. E. Apostolakis, and D. Okrent, “Computer-oriented approach to fault-tree

construction”, No. EPRI-NP-288, California University (1976).

69. Salem, S. L., J. S. Wu, and G. Apostolakis, "Decision table development and application to the

construction of fault trees", Nuclear Technology 42.1 (1979): 51-64.

70. Sheskin, T. J., “Markov chains and decision processes for engineers and managers”, CRC Press

(2016).

71. Siu, N. "Risk assessment for dynamic systems: an overview", Reliability Engineering &

System Safety 43.1 (1994): 43-73.

72. Siu, N., C. Acosta, and N. C. Rasmussen, “Physical dependencies in accident sequence

analysis”, Cambridge, MA: Nuclear Engineering Dept., Massachusetts Institute of

Technology (1989).

73. Smidts, C. "Probabilistic Reactor Dynamics—IV. An Example of Man/Machine Interaction",

Nuclear Science and Engineering 112.2 (1992): 114-126.

74. Smidts, C., and J. Devooght. "Probabilistic reactor dynamics—II: A Monte Carlo study of a

fast reactor transient", Nuclear Science and Engineering 111.3 (1992): 241-256.

75. Spek, J. A. W., “Cell mapping methods: modifications and extensions”, PhD Dissertation,

Eindhoven University of Technology, Netherlands (1994).

76. Stamatelatos, M., W. Vesely, J. Dugan, J. Fragola, J. Minarick, and J. Railsback, "Fault tree

handbook with aerospace applications", Office of Safety and Mission Assurance, NASA HQ,

(2002).

77. Taylor, Z., and Ranganathan, S., “Designing High Availability Systems: DFSS and Classical

Reliability Techniques with Practical Real Life Examples”, John Wiley & Sons (2013).

78. Tombuyses, B., and T. Aldemir, "Computational efficiency of the continuous cell-to-cell

mapping technique as a function of integration schemes", Reliability Engineering & System

Safety 58.3 (1997): 215-223.

Page 194: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-178

79. US Nuclear Regulatory Commission, "PRA Procedures Guide", NUREG/CR-2300,

Washington DC (1983).

80. US Nuclear Regulatory Commission, “Reactor Safety Study”, WASH-1400 (NUREG-

75/014), US Nuclear Regulatory Commission, Washington DC (1975).

81. Vesely, W. E., Goldberg, F. F., Roberts, N. H., & Haasl, D. F., “Fault tree handbook”,

NUREG/CR-0492, Nuclear Regulatory Commission, Washington DC (1981).

82. Wang, P., and T. Aldemir, "Some improvements in state/parameter estimation using the cell-

to-cell mapping technique", Nuclear Science and Engineering 147.1 (2004): 1-25.

83. Yang, J., and Aldemir, T., "An algorithm for the computationally efficient deductive

implementation of the Markov/Cell-to-Cell-Mapping Technique for risk significant scenario

identification", Reliability Engineering & System Safety 145 (2016): 1-8.

84. Yau, M., "Dynamic flowgraph methodology for the analysis of software-based controlled

systems", PhD Dissertation, University of California, Los Angeles (1997).

85. Yau, M., G. Apostolakis and S. Guarro, “The use of prime implicants in dependability analysis

of software controlled systems”, Reliability Engineering and System Safety, 62, 23-32 (1998).

86. Yau, M., M. Motamed and S. Guarro, “Assessment and Integration of Software Risk within

PRA”, International Journal of Performability Engineering, 3 (2007): 369-378.

87. Yau, M., S. Guarro, and G. Apostolakis, "Demonstration of the dynamic flowgraph

methodology using the Titan II space launch vehicle digital flight control system", Reliability

Engineering & System Safety 49.3 (1995): 335-353.

88. Zeliang, C., O. P. Singh, and P. Munshi, "Uncertainty evaluation of reliability of shutdown

system of a medium size fast breeder reactor", Nuclear Engineering and Design 308 (2016):

283-296.

89. Zio, E., "Integrated deterministic and probabilistic safety assessment: concepts, challenges,

research directions", Nuclear Engineering and Design 280 (2014): 413-419.

90. Zio, E., and N. Pedroni, "Building confidence in the reliability assessment of thermal-

hydraulic passive systems", Reliability Engineering & System Safety 94.2 (2009): 268-281.

Page 195: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-

APPENDIX I

Multi-state modelling with FT and Markov model

To demonstrate the capability of FT and Markov model in modeling multi-state

component/system, consider a simple non-repairable system consisting of two components

connected in parallel. Assume that the system is fully operational initially and is represented by

the system state 𝑆0 (both components in normal state) at 𝑡 = 0. Two distinct cases are considered

for analysis and illustration purpose.

1. Components with one failure modes (binary);

2. Components with two failure modes (multi-state).

1. Components with one failure modes

It is considered that each component is independent and has only one failure mode i.e., normal or

failed. The state transition probabilities are 𝜆𝐴 and 𝜆𝐵 for components A and B respectively. The

Markov state transition diagram for the redundant system is given in Figure A-I-1.

Figure A-I-1: Second-order failures Markov state transition diagram

Figure A-I-1 can be characterized by a set of ODEs:

𝑑𝑆0

𝑑𝑡= −(𝜆𝐴 + 𝜆𝐵)𝑆0

𝑑𝑆1

𝑑𝑡= 𝜆𝐴𝑆0 − 𝜆𝐵𝑆1

𝑑𝑆2

𝑑𝑡= 𝜆𝐵𝑆0 − 𝜆𝐴𝑆2

𝑑𝑆3

𝑑𝑡= 𝜆𝐵𝑆1 + 𝜆𝐴𝑆2

(A-I-1)

Solving the set of ODEs with the initial conditions 𝑆0 = 1 and 𝑆1 = 𝑆2 = 𝑆3 = 0, we obtain the

following state probabilities.

Page 196: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-i

𝑆0(𝑡) = 𝑒−(𝜆𝐴+𝜆𝐵)𝑡

𝑆1(𝑡) = 𝑒−𝜆𝐵 𝑡(1 − 𝑒−𝜆𝐴 𝑡)

𝑆2(𝑡) = 𝑒−𝜆𝐴 𝑡(1 − 𝑒−𝜆𝐵 𝑡)

𝑆3(𝑡) = (1 − 𝑒−𝜆𝐴 𝑡). (1 − 𝑒−𝜆𝐵 𝑡) (A-I-2)

Since a system failure occurs when it occupies state 𝑆3 (both component failure), the system failure

probability via Markov model is given by:

𝑆3(𝑡) = (1 − 𝑒−𝜆𝐴𝑡). (1 − 𝑒−𝜆𝐵𝑡) (A-I-3)

The system failure probability (top event) predicted by the FT technique for the redundant system

is given by an AND gate as follow:

Failure probability of component A, 𝑃𝐴 = (1 − 𝑒−𝜆𝐴𝑡) (A-I-4)

Failure probability of component B, 𝑃𝐵 = (1 − 𝑒−𝜆𝐵𝑡) (A-I-5)

𝑃(𝑇𝑜𝑝𝐸𝑣𝑒𝑛𝑡) = 𝑃𝐴 ∧𝑃𝐵 = (1 − 𝑒−𝜆𝐴𝑡). (1 − 𝑒−𝜆𝐵𝑡) (A-I-6)

Clearly, the top event (failure probability) predicted by FT and Markov model are the same.

2. Components with two failure modes

For the second case, components A and B are considered to have two failure modes 𝜆1𝐴, 𝜆2

𝐴, 𝜆1𝐵 and

𝜆2𝐵 respectively with the same system configuration as the first case. The state transition diagram

is shown in Figure A-I-1.

Figure A-I-2: Markov state transition diagram for second case

Page 197: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-ii

The set of ODEs associated with the state transition diagram are:

𝑑𝑆0(𝑡)

𝑑𝑡= −(𝜆1

𝐴 + 𝜆2𝐴 + 𝜆1

𝐵 + 𝜆2𝐵)𝑆0(𝑡)

𝑑𝑆1(𝑡)

𝑑𝑡= 𝜆1

𝐵𝑆0(𝑡) − (𝜆1𝐴 + 𝜆2

𝐴)𝑆1(𝑡)

𝑑𝑆2(𝑡)

𝑑𝑡= 𝜆2

𝐵𝑆0(𝑡) − (𝜆1𝐴 + 𝜆2

𝐴)𝑆2(𝑡)

𝑑𝑆3(𝑡)

𝑑𝑡= 𝜆1

𝐴𝑆0(𝑡) − (𝜆1𝐵 + 𝜆2

𝐵)𝑆3(𝑡)

𝑑𝑆4(𝑡)

𝑑𝑡= 𝜆2

𝐴𝑆0(𝑡) − (𝜆1𝐵 + 𝜆2

𝐵)𝑆4(𝑡)

𝑑𝑆5(𝑡)

𝑑𝑡= 𝜆1

𝐴𝑆1(𝑡) + 𝜆1𝐵𝑆3(𝑡)

𝑑𝑆6(𝑡)

𝑑𝑡= 𝜆2

𝐴𝑆1(𝑡) + 𝜆1𝐵𝑆4(𝑡)

𝑑𝑆7(𝑡)

𝑑𝑡= 𝜆1

𝐴𝑆2(𝑡) + 𝜆2𝐵𝑆3(𝑡)

𝑑𝑆8(𝑡)

𝑑𝑡= 𝜆2

𝐴𝑆2(𝑡) + 𝜆2𝐵𝑆4(𝑡)

(A-I-7)

Considering that the system is operational at 𝑡 = 0, initial condition is given as:

𝑆𝑛(0) = 0; 𝑓𝑜𝑟𝑛 ≠ 0 (A-I-8)

Where; 𝑛 = the number of system states

The resulting solution of the ODEs are:

𝑆0(𝑡) = 𝑒−(𝜆1𝐴+𝜆2

𝐴+𝜆1𝐵+𝜆2

𝐵)𝑡

𝑆1(𝑡) =𝜆1𝐵

𝜆1𝐵 + 𝜆2

𝐵 𝑒−(𝜆1

𝐴+𝜆2𝐴)𝑡 − 𝑒−(𝜆1

𝐴+𝜆2𝐴+𝜆1

𝐵+𝜆2𝐵)𝑡

𝑆2(𝑡) =𝜆2𝐵

𝜆1𝐵 + 𝜆2

𝐵 𝑒−(𝜆1

𝐴+𝜆2𝐴)𝑡 − 𝑒−(𝜆1

𝐴+𝜆2𝐴+𝜆1

𝐵+𝜆2𝐵)𝑡

𝑆3(𝑡) =𝜆1𝐴

𝜆1𝐴 + 𝜆2

𝐴 𝑒−(𝜆1

𝐵+𝜆2𝐵)𝑡 − 𝑒−(𝜆1

𝐴+𝜆2𝐴+𝜆1

𝐵+𝜆2𝐵)𝑡

𝑆4(𝑡) =𝜆2𝐴

𝜆1𝐴 + 𝜆2

𝐴 𝑒−(𝜆1

𝐵+𝜆2𝐵)𝑡 − 𝑒−(𝜆1

𝐴+𝜆2𝐴+𝜆1

𝐵+𝜆2𝐵)𝑡

𝑆5(𝑡) =𝜆1𝐴𝜆1

𝐵

𝜆1𝐴𝜆1

𝐵 + 𝜆1𝐴𝜆2

𝐵 + 𝜆2𝐴𝜆1

𝐵 + 𝜆2𝐴𝜆2

𝐵 1 + 𝑒−(𝜆1𝐴+𝜆2

𝐴+𝜆1𝐵+𝜆2

𝐵)𝑡 − 𝑒−(𝜆1𝐵+𝜆2

𝐵)𝑡 − 𝑒−(𝜆1𝐴+𝜆2

𝐴)𝑡

Page 198: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-iii

𝑆6(𝑡) =𝜆2𝐴𝜆1

𝐵

𝜆1𝐴𝜆1

𝐵 + 𝜆1𝐴𝜆2

𝐵 + 𝜆2𝐴𝜆1

𝐵 + 𝜆2𝐴𝜆2

𝐵 1 + 𝑒−(𝜆1𝐴+𝜆2

𝐴+𝜆1𝐵+𝜆2

𝐵)𝑡 − 𝑒−(𝜆1𝐵+𝜆2

𝐵)𝑡 − 𝑒−(𝜆1𝐴+𝜆2

𝐴)𝑡

𝑆7(𝑡) =𝜆1𝐴𝜆2

𝐵

𝜆1𝐴𝜆1

𝐵 + 𝜆1𝐴𝜆2

𝐵 + 𝜆2𝐴𝜆1

𝐵 + 𝜆2𝐴𝜆2

𝐵 1 + 𝑒−(𝜆1𝐴+𝜆2

𝐴+𝜆1𝐵+𝜆2

𝐵)𝑡 − 𝑒−(𝜆1𝐵+𝜆2

𝐵)𝑡 − 𝑒−(𝜆1𝐴+𝜆2

𝐴)𝑡

𝑆8(𝑡) =𝜆2𝐴𝜆2

𝐵

𝜆1𝐴𝜆1

𝐵+𝜆1𝐴𝜆2

𝐵+𝜆2𝐴𝜆1

𝐵+𝜆2𝐴𝜆2

𝐵 1 + 𝑒−(𝜆1𝐴+𝜆2

𝐴+𝜆1𝐵+𝜆2

𝐵)𝑡 − 𝑒−(𝜆1𝐵+𝜆2

𝐵)𝑡 − 𝑒−(𝜆1𝐴+𝜆2

𝐴)𝑡 (A-I-9)

The system unreliability is determined by summing all the second order failures.

��(𝑡) = 𝑆𝑛(𝑡)

8

𝑛=5

(A-I-10)

��(𝑡) =(1+𝑒

− 𝜆1𝐴+𝜆2

𝐴+𝜆1𝐵+𝜆2

𝐵 𝑡−𝑒

− 𝜆1𝐵+𝜆2

𝐵 𝑡−𝑒

− 𝜆1𝐴+𝜆2

𝐴 𝑡)

𝜆1𝐴𝜆1

𝐵+𝜆1𝐴𝜆2

𝐵+𝜆2𝐴𝜆1

𝐵+𝜆2𝐴𝜆2

𝐵 (𝜆1𝐴𝜆1

𝐵 + 𝜆2𝐴𝜆1

𝐵 + 𝜆1𝐴𝜆2

𝐵 + 𝜆2𝐴𝜆2

𝐵) (A-I-11)

Some differences between FTA and Markov model can be observed. To highlight one key

difference, consider the system state 𝑆8(𝑡), where both component A and B are in a failed state

with modes 𝜆2𝐴 and 𝜆2

𝐵 respectively. The individual failure probabilities of the components are:

𝑃𝐴2 = 1 − 𝑒−𝜆2𝐴𝑡 and 𝑃𝐵2 = 1 − 𝑒−𝜆2

𝐵𝑡 (A-I-12)

Therefore, system failure probability (top event) predicted by FT technique is computed as:

TopEvent = 𝑃𝐴2 ∧𝑃𝐵2 = 1 − 𝑒−𝜆2𝐴𝒕 . 1 − 𝑒−𝜆2

𝐵𝒕 (A-I-13)

Whereas, the failure probability of the top event computed from Markov model is given in

Equation (A-I-9). Clearly, it can be observed that the failure probability predicted by the FT and

Markov model are different. For illustration purpose, consider that the component A have two

failure modes with failure rates 𝜆1𝐴 = 0.005𝑝𝑒𝑟ℎ𝑟 and 𝜆2

𝐴 = 0.007𝑝𝑒𝑟ℎ𝑟, and component B

have two failure modes with failure rates of 𝜆1𝐵 = 0.005𝑝𝑒𝑟ℎ𝑟 and 𝜆2

𝐵 = 0.007𝑝𝑒𝑟ℎ𝑟

respectively. The two cases are simulated for comparison purpose.

• Components A and B with binary mode (normal and failed)

• Components A and B with multiple modes, and system failure occurs when A and B are in

𝜆2𝐴 and 𝜆2

𝐵 respectively.

Page 199: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-iv

It may be observed from Figure A-I-3 that the top event probabilities computed from the two

methods are different. The top event computed from Markov model is smaller than the one from

the FT technique (green and red curve). This is due to the fact that Markov model accounts for

competition between failure modes of a component, while the FT treats the two failure modes

independently. For the explicit mode consideration, top event from FT is simply given by an AND

gate between 𝜆2𝐴 and 𝜆2

𝐵, whereas in Markov model, 𝑆8(𝑡) represents the top event.

Figure A-I-3: Top event comparison between FT and Markov model

The asymptotic behavior of the system predicted by FT tends to ‘1’ as (𝑡 → ∞), whereas the

system state 𝑆8(𝑡) from the Markov model as (𝑡 → ∞) is given by:

𝑆8(𝑡 → ∞) =𝜆2𝐴𝜆2

𝐵

𝜆1𝐴𝜆1

𝐵+𝜆1𝐴𝜆2

𝐵+𝜆2𝐴𝜆1

𝐵+𝜆2𝐴𝜆2

𝐵 (A-I-14)

i.e., 𝑆8(𝑡 → ∞) = 0.3403 (limiting state probability), as opposed to ‘1’ predicted by the FT

technique. These properties of Markov model accounting for failure modes in an integrated fashion

provide a superior way to model components/systems with multiple failure modes or states.

Page 200: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-v

APPENDIX II

! Chireuding Zeliang/Major: Nuclear Engineering

! Faculty of Energy Systems and Nuclear Science: University of Ontario Institute of Technology,

Oshawa, Ontario, 2018

Program Benchmark_Liquid_Level_Control_System

! Declaration of variables

Real :: M(27,27), A(27,27), P(11,27), dt, t, q(27), sum

Integer :: i, j, k

! Initialization of variables

i=0

j=0

k=0

dt=1.

do i= 1, 27

do j= 1, 27

M(i,j)=0.

A(i,j)=0.

if (i.eq.j) then

A(i,j)=1. ! Identity matrix

end if

end do

end do

do k=1,11 ! The number of time steps

do i=1,27

P(k,i)=0.

q(i)=0. ! Initial transition matrix

end do

end do

P(1,1)=1. ! Initial state of the system

! Coefficient of the coupled Ordinary Differential Equations

Page 201: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-vi

M(1,1): − 0.0134 M(2,1): 0.0023 M(2,2): −0.0088

M(3,1): 0.0023 M(3,3): −0.0088 M(4,1): 0.0029

M(4,4): −0.0077 M(5,1): 0.00285 M(5,5): −0.0077

M(6,1): 0.0016 M(6,6): −0.0103 M(7,1): 0.00156

M(7,7): −0.0103 M(8,2): 0.0028 M(8,4): 0.0023

M(8,8): 0.0031 M(9,2): 0.0028 M(9,5): 0.0023

M(9,9): −0.0031 M(10,2): 0.0016 M(10,7): 0.0023

M(10,10): −0.0057 M(11,2): 0.0016 M(11,6): 0.0023

M(11,11): −0.0057 M(12,3): 0.0029 M(12,4): 0.0023

M(12,12): −0.0031 M(13,3): 0.0029 M(13,5): 0.0023

M(13,13): −0.0031 M(14,3): 0.00156 M(14,6): 0.0023

M(14,14): −0.0057 M(15,3): 0.00156 M(15,7): 0.0023

M(15,15): −0.0057 M(16,4): 0.00156 M(16,6): 0.00285

M(16,16): −0.0046 M(17,4): 0.00156 M(17,7): 0.00285

M(17,17): −0.0046 M(18,5): 0.00156 M(18,6): 0.00285

M(18,18): −0.0046 M(19,5): 0.00156 M(19,7): 0.00285

M(19,19): −0.0046 M(20,8): 0.00156 M(20,10): 0.00285

M(20,16): 0.0023 M(21,8): 0.00156 M(21,11): 0.00285

M(21,17): 0.0023 M(22,9): 0.00156 M(22,10): 0.00285

M(22,18): 0.0023 M(23,9): 0.00156 M(23,9): 0.0016

M(23,11): 0.00285 M(23,19): 0.0023 M(24,12): 0.0016

M(24,14): 0.00285 M(24,16): 0.0023 M(25,12): 0.0016

M(25,15): 0.00285 M(25,17): 0.0023 M(26,13): 0.0016

M(26,14): 0.00285 M(26,18): 0.0023 M(27,13): 0.0016

M(27,15): 0.00285 M(27,19): 0.0023

! Computation of transition state probabilities

do k=1,10

t=(k-1)*dt

do i=1,27

do j=1,27

q(i)= (A(i,j) + M(i,j)*dt) * P(k,j)

Page 202: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-vii

P(k+1, i)= P(k+1,i) + q(i)

end do

end do

end do

! Results

write(1,*)"Markov model of the benchmark system"

write(1,*)""

write(1,*)"Sort by time"

write(1,*)""

do k=1,5

t=(k-1)*dt

write(1,*)"t=",t

sum=0

do i= 1,27

sum= sum + P(k,i)

write(1,*)"P(",i,")=", P(k,i)

end do

write(1,*)"sum=",sum

write(1,*)""

end do

write(1,*)""

write(1,*)"Systematically arrange by system status"

write(1,*)""

do i=1,27

write(1,*)"n=", i

do k=1,5

write(1,*) P(k,i)

end do

write(1,*)""

end do

end

Page 203: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-viii

APPENDIX III

Timed Fault Tree generation from DFM Model

Figure A-III-1: Timed-fault tree for transfer gate G5

Page 204: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-ix

Figure A-III-2: Timed-fault tree for transfer

gate G14

Figure A-III-3: Timed-fault tree for transfer

gate G15

Page 205: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-x

Figure A-III-4: Timed-fault tree for transfer

gate G7

Figure A-III-5: Timed-fault tree for transfer

gate G8

Page 206: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xi

Figure A-III-6: Reduced timed-fault tree after consistency check

Page 207: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xii

APPENDIX IV

State-space Discretization and computation of cell-to-cell transition probabilities.

clear all;

clc;

% State-Space Discretization Algorithm

syms VR Vj1 Vj2 Vj3 Vr1 Vr2 Vr3 r Vr x(t) dt f f1 z J1 J2 J3

VR_L= -3; % Control space lower bound

VR_U= +3; % Control space upper bound

VR= [-3:+3]; % Overall Control Space

VR= [-3:2:+3]; %Discretized control regions: analyst choice

for VR= -3:2:+3

if (VR<=-3)

Vr1_L=VR;

elseif (VR<=-1)

Vr1_U=VR;

Vr2_L=Vr1_U;

elseif (Vr2_L<VR<1)

Vr2_U=VR;

Vr3_L=Vr2_U;

elseif (VR>=+3)

Vr3_U=VR;

end

end

r=[1 2 3]; %Number of control region

Vr1= [-3 -1]; % Space for control region-1

Vr2= [-1 +1]; % Space for control region-2

Vr3= [+1 +3]; % Space for control region-3

r1= range(Vr1); %Range of control region-1

cw1= range(Vr1)/3; %computational cell width in region-1

r2= range(Vr2); %Range of control region-2

cw2= range(Vr2)/3; %computational cell width in region-2

r3= range(Vr3); %Range of control region-3

cw3= range(Vr3)/3; %computational cell width in region-3

Page 208: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xiii

% Discretization scheme for Control region-1

a_bar= -3; % Lower bound of control space Vr1

alpha_1= -1; % Upper bound of control space Vr1

J1=3; % Number of sub-cells in control space Vr1

for j=1:J1

delta_x1= ((alpha_1-a_bar)/J1); %sub-cell size/volume

a(j)= a_bar + (j-1)*delta_x1; % Lower bound of the sub-cells

b(j)= a_bar + j*delta_x1; % Upper bound of the sub-cells

dp1(:,j)= ((a(:,j)+b(:,j))/2); %departure points from the sub-cells of control region-1

Vd(j)=-3;

end

% Discretization scheme for Control region-2

alpha_1= -1; % Lower bound of control space Vr2

alpha_2= +1; % Upper bound of control space Vr2

J2=3; % Number of sub-cells in control space Vr2

for j=1:J2

delta_x2= ((alpha_2-alpha_1)/J2); %sub-cell size/volume

a(j)= alpha_1 + (j-1)*delta_x2; % Lower bound of the sub-cells

b(j)= alpha_1 + j*delta_x2; % Upper bound of the sub-cells

dp2(:,j)= ((a(:,j)+b(:,j))/2); %departure points from the sub-cells of control region-2

end

% Discretization scheme for Control region-3

alpha_2= +1; % Lower bound of control space Vr3

b_bar= +3; % Upper bound of control space Vr3

J3=3; % Number of sub-cells in control space Vr3

for j=1:J3

delta_x3= ((b_bar-alpha_2)/J3); %sub-cell size/volume

a(j)= alpha_2 + (j-1)*delta_x3; % Lower bound of the sub-cells

b(j)= alpha_2 + j*delta_x3; % Upper bound of the sub-cells

dp3(:, j)= ((a(:,j)+b(:,j))/2); %departure points from the sub-cells of control region-3

Vo(j)=+3;

end

Page 209: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xiv

DP= [dp1; dp2; dp3]; %Departure points from all the sub-cells of the three regions

x(t)=DP;

%System dynamics is defined by (dx/dt)=f

x1_bar= 1; %flowrate from unit-1

x2_bar= 1; %flowrate from unit-2

x3_bar= 1; %flowrate from unit-3

N=8; % component state combination

for n=1:8;

for dt=1 %the time step

f= (x1_bar + x2_bar - x3_bar)*dt; %Liquid level evolution as a function of N

if n==1 % system at component state combination n=1, 4 and 6

z1=f; %z1 is the "dx" i.e., the small increment in level

y1=[x(t)+z1]; % y1= x(t+dt) is the liquid level in the next time step

y11= vpa(y1);

else if n==2 % system state at n=2

z2=2*f; %The net flowrate at n=2

y2=[x(t)+z2]; % y2=x(t+dt) is the liquid level in the next time step

y22= vpa(y2); %Fraction to decimal

elseif n==3 % system state at n=3, 5 and 8

z3=0;

y3=[x(t)+z3];

y33=vpa(y3);

elseif n==7 % system state at n=7

z4=-f;

y4=[x(t)+z4];

y44=vpa(y4);

end

end

end

end

Page 210: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xv

p=3; %number of rows and column of y11, y22, y33 and y4

B=zeros(5,5);

B(1,1)=-5;

for i=1:5

B(5,i)=+5;

end

for i=1:p

for j=1:p

B(i+1,j+1)=y11(i,j);

end

end

J= 5; %number of discretized space

h=5; % number of groups/control range

k(J,h)=zeros; % k is a group matrix

% Each row and column or elements of matrix 'k' represents the number of arrival trajectories in j from j'

for j=1:J %Number of columns of g(j/j',n',dt)

for i=1:J %Number of rows g(j/j',n',dt)

if B(i,j)<-3

k(i,1)=k(i,1)+1;

end

if B(i,j)>=-3 && B(i,j)<=-1

k(i,2)=k(i,2)+1;

end

if B(i,j)>-1 && B(i,j)<=+1

k(i,3)=k(i,3)+1;

end

if B(i,j)>+1 && B(i,j)<=+3

k(i,4)=k(i,4)+1;

end

if B(i,j)>+3

k(i,5)=k(i,5)+1;

end

end

end

Page 211: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xvi

APPENDIX V

Sample Results: State Transition Probabilities of the Passive Isolation Condenser Systems

Transition State Probabilities for 𝑘 = 5:

𝑃1 = 0.997 𝑃2 = 1.37E-03 𝑃3 = 1.59E-07

𝑃4 = 1.75E-10 𝑃5 = 2.621E-05 𝑃6 = 2.88E-08

𝑃7 = 3.36E-12 𝑃8 = 2.77E-15 𝑃9 = 2.62E-06

𝑃10 = 2.87E-09 𝑃11 = 3.36E-13 𝑃12 = 2.77E-16

𝑃13 = 0.00 𝑃14 = 4.44E-18 𝑃15 = 5.19E-22

𝑃16 = 5.689E-25 𝑃17 = 8.51E-20 𝑃18 = 9.33E-23

𝑃19 = 1.09E-26 𝑃20 = 6.73E-30 𝑃21 = 8.51E-21

𝑃22 = 9.33E-24 𝑃23 = 1.09E-27 𝑃24 = 6.73E-31

𝑃25 = 6.02E-04 𝑃26 =6.61E-07 𝑃27 = 7.73E-11

𝑃28 = 6.36E-14 𝑃29 = 1.27E-08 𝑃30 = 1.04E-11

𝑃31 = 1.22E-15 𝑃32 = 6.68E-19 𝑃33 = 1.27E-09

𝑃34 = 1.04E-12 𝑃35 = 1.22E-16 𝑃36 = 6.68E-20

𝑃37 = 6.02E-04 𝑃38 = 6.61E-07 𝑃39 = 7.73E-11

𝑃40 = 6.36E-14 𝑃41 = 1.27E-08 𝑃42 = 1.04E-11

𝑃43 = 1.22E-15 𝑃44 = 6.68E-19 𝑃45 = 1.27E-09

𝑃46 = 1.04E-12 𝑃47 = 1.22E-16 𝑃48 = 6.68E-20

𝑃49 = 1.95E-18 𝑃50 = 2.14E-21 𝑃51 = 2.51E-25

𝑃52 = 1.55E-28 𝑃53 = 4.11E-23 𝑃54 = 2.54E-26

𝑃55 = 2.97E-30 𝑃56 = 8.67E-34 𝑃57 = 4.11E-24

𝑃58 = 2.54E-27 𝑃59 = 2.97E-31 𝑃60 = 8.67E-35

𝑃61 = 2.91E-07 𝑃62 = 2.39E-10 𝑃63 = 2.80E-14

𝑃64 = 1.54E-17 𝑃65 = 4.59E-12 𝑃66 = 2.52E-15

𝑃67 = 2.95E-19 𝑃68 = 8.07E-23 𝑃69 = 4.59E-13

𝑃70 = 2.52E-16 𝑃71 = 2.95E-20 𝑃72 = 8.07E-24

𝑆𝑢𝑚 = 1.00

Page 212: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xvii

APPENDIX VI

Dynamic Probabilistic Safety

Assessment using Dynamic

Flowgraph Method and Markov-

Cell-to-cell Mapping Technique

Chireuding Zeliang

MASc Candidate

Supervisor:

Prof. Akira Tokuhiro

(Prof. Lixuan Lu)

24th July 2018

Acknowledgements

• My sincere gratitude to Prof. Akira Tokuhiro

• My appreciation to the Committee members:

Prof. Glenn Harvel and Dr. Salam K. Ali

• Dr. Hadid Subki, Nuclear Power Technology Development Section,

International Atomic Energy Agency (IAEA)

2

Objectives

• Demonstrate the advantages and applicability of dynamic PSA

methods over classical techniques

• Formulate the coupling of Markov-CCMT model for dynamic PSA

• Illustrate generation of timed-fault trees from DFM model for its

integration into an existing classical PSA model

• Development of a novel approach for integrated reliability

assessment of passive safety system in iPWR-type Small Modular

Reactors (IAEA-UOIT Coordinated Research Project)

3

Introduction

• Classical Probabilistic Risk Assessment (PRA)

➢ Event/Fault Tree techniques

➢Classical Markov chain

• Dynamic PRA

➢Dynamic Flowgraph Method

➢Coupled Markov/Cell-to-cell Mapping Technique

• Application and demonstration of methodologies

• Development of an integrated approach for Dynamic Reliability

Assessment of passive safety systems in iPWR-type SMRs

4

NUREG-6901: “Current State of Reliability Modeling Methodologies for Digital Systems and Their Acceptance Criteria for Nuclear Power Plant Assessments” (2016)

Classical PSA techniques

• Well-established techniques, and a systematic and

comprehensive approach for Risk Assessment

• Fault/Event Tree Technique (WASH-1400, NUREG-75/014, 1975)

• Classical Markov chain (Norris, 1997)

• Static logic based technique

• Fault trees are series of Boolean Algebraic equations to

determine the failure probability of an undesired event

5

• WASH-1400, NUREG-75/014, “Reactor Safety Study: An assessment of accident risks in U.S. Commercial Nuclear power plants”, US Nuclear Regulatory Commission, 1975.

• Norris, J. R., “Continuous-time Markov chains I and II”, 1997.

Limitations of Classical PSA

• Classical PRA techniques lacks the ability to account for

system dynamics

• Lacks the ability to capture dynamic interactions

– Component failure modes and probability as a function of state

variables (e.g., interactions of control units through liquid level)

• Classical techniques are neither developed nor intended to

capture time element

• States ordering in ET/FT are fixed/no particular ordering index

• FT (binary logic) has limited capability in multi-state modelling

6

Page 213: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xviii

Dynamic PSA

• Dynamic PRA techniques enables one to couple system model with risk analysis code (as early as 1986)

• Simultaneously accounting for system time-dependent phenomenological model and its stochastic behavior

• Explicit consideration and treatment of time-element

• A better treatment of dynamic interactions and scenario development

➢ For systems with multiple top events, the final state is dependent on magnitude of state variables, order and timing of failure events

• Allows for systematic and more complete coverage of possible failure scenarios

• State-space explosion phenomenon, highly time-dependent, modelling complexity and computational time requirements

7

Dynamic Flowgraph Method (DFM)

• An integrated approach to modelling system logical and time-

dependent behavior

• Multi-valued logic model

• Dynamic behaviors are represented as a series of discrete state

transitions

• A single model can analyze arbitrary number of possible system

conditions (i.e., any top event of interest)

• A three step process:

1. Model development

2. Model analysis (deductive/inductive)

3. Quantification of the prime implicants.

8

Prime Implicants

• A prime implicant is a conjunction of basic events that is sufficient to cause a top event, but does not contain any shorter conjunction of events that is sufficient to cause the top event

• Analogous to minimal cut-set in FT, but is time-stamped (e.g., Ai @t=-1: component ‘A’ is in state “i” at time “t=-1”)

<Variable A=2 @t=-1 AND Variable A=3 @t=-2>

• Complete base is the set of prime implicants logically equivalent to the FT TOP event function

• Complete base is obtained by evaluation of transition table using algebraic laws including:

• Absorption, Merging, Reduction and Absorption-Merging law (Yau, 1997)

• Generalized Consensus Method (Garribba et al., 1985)

9

Timed FT generation from DFM model

• Integration of results obtained from DFM model to an existing

classical plant PRA model

• Sequences of static FTs at different time steps representing

evolution of logical combinations of events leading to a top event

• Consistency Rules:

– Physical consistency: A process variable with different state cannot

occur in the same time step. e.g., valve ‘A’ open AND valve ‘A’ closed @t= -1

– Dynamic consistency: A process variable can change its states in a

certain direction or by amount (dictated by system dynamics) in a single

time step. e.g., liquid level variation from 0% to 90% in a single time step

10

Coupled Markov-CCMT Model

System Description

System Analysis

Analysis of Type-I

Interactions

Analysis of Type-II

Interactions

Dynamic and Control

Laws

Cell-to-cell Mapping

Technique

Finite States

Description

Markov Model

System Modelling

(Deterministic)

Stochastic System

Model (Probabilistic)

Integrated Deterministic and

Probabilistic Analysis

11

Page 214: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xix

The Algorithm

• The probability 𝑷𝒏,𝒋 𝒕 of finding a state variable in cell 𝒋 at time

= ∆ for a given component state combination ′𝒏′ can be

recursively found:

𝑷𝒏,𝒋 𝒌+ 𝟏 = 𝑔 𝑗 𝑗′⁄ , 𝑛′, ∆𝑡 . ℎ 𝑛 𝑛′⁄ , 𝑗′ → 𝑗, ∆𝑡 . 𝑃𝑛′.𝑗′(𝑘)

𝐽

𝑗′=1

𝑁

𝑛′=1

𝑷𝒏,𝒋 𝒌 + 𝟏 = 𝒏,𝒋𝒏′,𝒋′

. 𝑃𝑛′.𝑗′(𝑘)

𝐽

𝑗′=1

𝑁

𝑛′=1

𝒏,𝒋𝒏′,𝒋′

= 𝑔 𝑗 𝑗′⁄ , 𝑛′, ∆𝑡 . ℎ 𝑛 𝑛′⁄ , 𝑗′ → 𝑗, ∆𝑡

13

′, ′, ∆ ⁄ : Conditional probability of state variables in cell 𝑗 at time (𝑘 + 1)∆𝑡 given that it was in cell 𝑗′ at time 𝑡 ′⁄ , ′ → , ∆ : Conditional probability of component state combination in state 𝑛 at time (𝑘 +1)∆𝑡 given that it was in state 𝑛′ at time 𝑘∆𝑡, and the state variables move from cell 𝑗′ to cell 𝑗 during 𝑘∆𝑡 ≤ 𝑡 ≤ (𝑘 + 1)∆𝑡}

Application of Methodologies

1. Fault Tree

2. Event Tree

3. Classical Markov model

4. Dynamic Flowgraph Method

5. Markov-CCMT model

14

The Benchmark System (BS)

Liquid Level

( )

Control Unit State

Unit-1 Unit-2 Unit-3

𝑥 > 𝑏 - - -

𝑏1 < 𝑥 OFF OFF ON

𝑎1 ≤ 𝑥 ≤ 𝑏1 ON OFF ON

𝑥 < 𝑎1 ON ON OFF

𝑥 < 𝑎 - - -

Control Laws

Unit discrete states (i)

• ON/OFF

• Failed-open

• Failed-closed

15

Fault Tree Model: Binary

16

Top Event

MCS 𝟏 𝟐 + 𝟑

Failure Probability 𝟑. 𝟏 𝟏𝟏 × 𝟏𝟎−𝟑

Ref: Electric Power Research Institute, “CAFTA Fault Tree Analysis System Version 6.0”, EPRI Inc., NC, 2013.

Fault Tree Model: Multiple Modes

17

Top Event: Overflow

MCS 𝟏 𝟐 + 𝟏 𝟑 + 𝟐 𝟑

Failure Probability 𝟏. 𝟒 × 𝟏𝟎−𝟎

Top Event: Dryout

MCS 𝟏 𝟐 𝟑

Failure Probability 𝟏. 𝟎𝟏 𝟐 × 𝟏𝟎−𝟎

Event Tree Model

18

Ref: Electric Power Research Institute, “CAFTA Fault Tree Analysis System Version 6.0”, EPRI Inc., NC, 2013.

System Overflow

𝟐. 𝟏 × 𝟏𝟎−𝟎

System Dryout

𝟑. 𝟎 × 𝟏𝟎−𝟎

Note: Rounded to two decimal point for simplicity

Page 215: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xx

Classical Markov model

19

Markov chain

• Enables one to model a sequence of random variables which

corresponds to discrete system states

• System state at any point in time is dependent only on the previous

state, i.e., conditional state probabilities

ℙ( 𝒏+𝟏 = 𝒋 𝒏 = 𝒊, 𝒏−𝟏 = 𝒊𝒏−𝟏, … , 𝟎 = 𝒊𝟎 = ℙ( 𝒏+𝟏 = 𝒋 𝒏 = 𝒊

• System dynamics represented by a set of coupled linear differential

equations (Chapman-Kolmogorov)

𝑑𝑃𝑖(𝑡)

𝑑𝑡= − 𝑎𝑖𝑗𝑃𝑖(𝑡)

𝑛

𝑗=1𝑗 𝑖

+ 𝑎𝑗𝑖𝑃𝑗(𝑡)

𝑛

𝑗=1𝑗 𝑖

• State probabilities can be determined by solving the set of coupled

ordinary differential equations

20

Enumeration of system states

Individual Unit States System States (𝒏)Unit-1 Unit-2 Unit-3

𝑛1 = 0 𝑛2 = 0 𝑛3 = 0 𝑃0(𝑡)

𝑛1 = 1 𝑛2 = 0 𝑛3 = 0 𝑃1(𝑡)

: : : :

: : : :

𝑛1 = 2 𝑛2 = 2 𝑛3 = 2 𝑃26(𝑡)

= (𝑷 𝒊 𝒇 𝒊 )(𝒏 𝒇 𝒏𝒊𝒕 )

= 𝟐

21

Markov state transition diagram

22

Layer-2Layer-1 Layer-3Layer-0

(Zero unit failure) (One unit failure) (Two unit failure) (Three unit failure)

Transition State Probabilities

𝑑𝑃(𝑡)

𝑑𝑡= 𝐴.𝑃(𝑡)

𝑃(𝑡 + ∆𝑡) = (𝐼 + 𝐴.∆𝑡).𝑃(𝑡)

Initial conditions:

𝑃𝑛 𝑡 = 0 = 1; 𝑓𝑜𝑟𝑛 = 00; 𝑓𝑜𝑟𝑛 ≠ 0

23

Sample Results 𝑘 = 3

System overflow: 8.6805 × 10−05

System dryout: 6.1155 × 10−08

DFM Model of the Benchmark System

24

Discrete states

Transition table captures system dynamics

Process variable node

Ref: DYMONDA 7.0, ASCA Inc., USA, 2013

Page 216: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xxi

Transition and Decision Table

25

State Variable DiscretizationDiscrete Unit States

Transition table: System dynamics

26

Prime Implicants for ‘system overflow’

# Prime Implicants Time Logic Probability

1. Tank Liquid Level was at +3 meters @t= -1 -

2. Tank Liquid Level was between +1 and +3 meters

Unit-1 stucked-open

Unit-3 stucked-closed

@t= -1 AND 6.5232 × 10−6

@t= -1 AND

@t= -1

3. Tank Liquid Level was between +1 and +3 meters

Unit-2 stucked-open

Unit-3 stucked-closed

@t= -1 AND 4.4643 × 10−6

@t= -1 AND

@t= -1

4. Tank Liquid Level was between +1 and +3 meters

Unit-1 stucked-open

Unit-2 stucked-open

@t= -1 AND 3.5673 × 10−6

@t= -1 AND

@t= -1

Top Event Probability 𝟏. 𝟒 𝟒 × 𝟏𝟎−

PIs for ‘system overflow’: 2 backtracking depth

# Prime Implicants Time Logic

1. Tank Liquid Level was between -1 and +1 meters

Unit-1 is Normal

Unit-2 Stucked-High

Unit-3 is Normal

Unit-3 Stucked-Low

@t= -2 AND

@t= -2 AND

@t= -2 AND

@t= -2 AND

@t= -1

2. Tank Liquid Level was between +1 and +3 meters

Unit-1 Stucked-High

Unit-3 is Normal

Unit-3 Stucked-Low

@t= -2 AND

@t= -2 AND

@t= -2 AND

@t= -1

27

State ordering/sequence

State variable evolution depicting system dynamics

Scenario history

Generation of timed fault tree

28

Final MCS via Backtracking process

For 5 discretized intervals

# Minimal cut-sets Logic

1. TL= +2 @t= –1 OR

2. (TL= +1 AND U1-S= +1 AND U2-S= +1) @t= –1 OR

3. (TL= +1 AND U1-S= +1 AND U3-S= –1) @t= –1 OR

4. (TL= +1 AND U2-S= +1 AND U3-S= –1) @t= –1

29

Markov-CCMT Model of the BS

• State space discretized into 5 cells, (𝐽 = 5)

➢3 computational cells

➢2 sink cells

• Computational cells are further discretized into 3 sub-cells

• Total number of system states, (𝑁 = 27)

• 𝑁 × 𝐽 = 135; 𝑞𝑛,𝑗𝑛′,𝑗′

𝑘∆𝑡 = 135 × 135square matrix

• Sink cells:

TopEvent = 𝑗 = 5; 𝑂𝑣𝑒𝑟𝑓𝑙𝑜𝑤𝑗 = 1; 𝐷𝑟𝑦𝑜𝑢𝑡

30

Page 217: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xxii

Predicted system state probabilities

Time step 𝑷𝒏,𝟏 𝒌∆𝒕 𝑷𝒏, 𝒌∆𝒕 𝑷 𝒌∆𝒕

𝒌 = 𝟏 0 0 0

𝒌 = 𝟐 0 1.019 × 10−08 1.019 × 10−08

𝒌 = 𝟑 1.857 × 10−07 9.173 × 10−05 9.192 × 10−05

𝒌 = 𝟒 2.1496 × 10−06 5.114 × 10−04 5.135 × 10−04

31

𝒏 = System states/configuration𝒋 = Number of discretized state variables = Total failure probability (Overflow + Dryout)

IAEA-UOIT Coordinated Research Project

(iPWR-type Small Modular Reactors)

32

IAEA-UOIT Coordinated Research Project

• Presentations made at two (2) IAEA technical meetings:

1. First Technical Meeting (30th Oct.- 2nd Nov. 2017)

- Passive Safety Features in iPWR-type SMRs:

- Overview and Planning of Technical Contract No. 21051

2. Second Technical Meeting (7-10th May 2018)

- Planning and Progress of the Research Project

33

Research proposal made: 17th October 2016

Technical contract awarded: 13th October 2017

Title: Application of DFM and FTA for Reliability and Risk Assessments of Passive Safety System in an Integral-PWR type SMRs

Preliminary CRP Roadmap

2017 2018 2019 2020

2016

• Research proposal awarded

• Review of SMR technology

• Review of Passive Safety Systems

• Review of Passive Safety Systems in iPWR SMRs

• Selection of Benchmark PSS to be analyzed

• Selection of methodology

• FMEA• Development of

Stochastic and Physical model

• Classical approach to system modelling

• Coupling of stochastic-physical model

• Reference case simulation

• Initial Drafting of IAEA-UOIT project

• Integration of draft from all participating countries

• Final drafting of the Technical Document

17th October:

Research Proposal

18th July

34

System Selection

• System identification• System mission • Success/failure criteria

Physical Model

• Development of model (best estimate code)• Reference case simulation• Validation of model

Classical approach

• Reliability analysis using Fault Tree• Sensitivity and Uncertainty analysis

Stochastic model

• Finite state machine• Time-dependent Reliability analysis using classical Markov model

Dynamic approach

• Coupling of Physical model with Stochastic model • Integrated Reliability and Uncertainty analysis: Markov-CCMT approach

System characteristic

• Operational characteristics • System parameter identification

• Identification of failure modes• Failure data base preparation

The Benchmark Isolation Condenser System

36

• Benchmark system is selected in alignment with other participating countries

• Enables one to validate the model and compare the predicted system reliability

• Some generic experimental data available (Argentina and India)

Page 218: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xxiii

Dynamic PRA for SMRs

• Dynamic PRA can be more suitable for integral

pressurized water reactor type SMRs:

• Simplified Design

• Reduced number of components (e.g., No reactor

coolant pumps)

• Reduced number of accident initiators (e.g., No large

brake loss-of-coolant accident)

• Less reliance on active power source

37

ICS Mission

• Primary objectives of the ICS includes:

▪ Reactor pressure vessel depressurization without the loss of primary coolant inventory

▪ Remove sensible and decay heat from the reactor for 72 hours (mission time)

▪ Maintain fuel peak cladding temperature within design limits (1200 oC typically for PWR).

• Advantages of implementing ICS includes:

▪ Prevents unnecessary activation of safety relief valves (SRVs)

▪ Eliminate/mitigate potential LOCA that may result from SRVs failure

▪ Potential elimination/delay of high pressure coolant injection system operation

38

Benchmark System Characterization

39

Condensate Drain Unit: Sample

State Transition Diagram

Unit state ordering

40

Predicted State Transition Probabilities

• The analysis shows that 𝑃1 𝑡 has the highest state transition

probability. This implies that the ICS is mostly likely to fail due

to an envelope failure/pressure boundary being compromised

• 𝑃1 𝑡 is followed by 𝑃24 𝑡 and 𝑃36 𝑡 which are the vent and

drain unit failure respectively. This can be due to the

redundancy provided for the vent and drain unit

• 𝑃36 𝑡 being lower than 𝑃24 𝑡 is due to the diversity provided

in the drain valves as compared to the identical vent valves.

41

CNSC, REGDOC- 2.5.2: Safety system failure probability requirement: < 𝟏𝟎−𝟎𝟑 (2014)

Conclusion

• Dynamic techniques can provide useful insight and significant information in risk assessment compared to classical techniques

• Detail analysis of the benchmark system shows that dynamic techniques can capture:

• System dynamics

• Time element

• State ordering

• A more realistic treatment of system/components with multiple top events and failure modes respectively

• Markov-CCMT model provide a more rigorous treatment of time element (real-time) or tightly coupled

• DFM have the advantage of providing a scenario history and probabilistic accident sequence evolution

42

Page 219: Dynamic Accident Sequence Analysis using Dynamic Flowgraph

A-xxiv

Conclusions (cont’)

• The issue of state space explosion can be systematically treated by:

➢ System state merging

➢ Hybrid approach (small FT and Large Markov model)

➢ Re-arranging and solving the matrix in Canonical form

• Classical approach yields conservative results, and is suitable for

identifying logical correctness and establishing conservative

relationships between top and basic events

• The integrated approach presented in this research provide a

framework for dynamic risk assessment of passive safety systems

that has never been employed.

43

Future Works

• Development of ICS physical model (deterministic model) using

best-estimate system code RELAP5

• Mechanized coupling of stochastic and deterministic system model

through a user-friendly computer code.

e.g., coupling of stochastic model (Fortran95) and physical model

(RELAP5) in Python code

• Integrated modelling of automatic depressurization system (ADS)

and ICS to capture the dynamic interactions, and their influence on

demand frequency and predicted failure probability of the two

systems

44

Thank You

“Towards an Integrated Risk Assessment”

45

@C.Zeliang, 2018